content
stringlengths 1
15.9M
|
---|
\section{Introduction}
The search of Lagrangian models, describing spinning particles,
has a long story. Most popular approach in this direction is
the formulation of the Lagrangian on the space-time,
extended by the anticommuting variables,
which upon quantization provide the system with the nontrivial spin.
This approach is closely related with the supersymmetric field theories
and is essential for the formulation of superparticle systems.
There is another, pure bosonic, approach in which the Lagrangian
is formulated on the direct product of the
initial space-time by some orbit of Poincar\'e group.
The actions of this sort are, in fact, the particle counterparts
of Born-Infeld type systems. This approach seems to be interesting
due its visible relation with orbit method of Kirillov-Konstant-Souriau
\cite{souriau}. Most developed investigation of such systems has been
presented in Refs. \cite{tomsk}.
Notice that the bosonic approach is the only correct
in $(2+1)$-dimensional systems, due to anyonic nature of planar particle.
The bosonic approach admits an aesthetically attractive modification,
where the additional bosonic variables are encoded
in the dependence of the Lagrangian on higher-order derivatives.
These systems seem to be interesting not only for their clear geometrical
meaning. The higher-derivative parts in their
Lagrangians can be generated by some
field-theoretical mechanism, e.g. arising as quantum corrections.
Investigation of such sort of particle systems
became popular after remarkable work of
Polyakov \cite{polac}, where he show, that evaluation of
the effective action of $CP^1$ model minimally coupled to the
Chern-Simons field for the charged solitonic excitation
results in the action
\begin{equation}
{\cal S}_{eff} =\int (m + {\pi\over 2\theta}K_2)|d{\bf x}|,
\label{polac}\end{equation}
where $K_2$ is world-line's torsion and $\theta$ is field coupling
strength.
Later it was found, that this system describes anyonic analog of
Majorana field equations \cite{3d}.
Due to further studies it became a part of physical folklore
that relativistic systems can get nontrivial spin,
if one adds to the Lagrangians the higher-derivatives terms
(more precisely, the terms, depending on
reparametrization invariants (extrinsic curvatures) of world-line).
Some significant observations were done in connection with this
subject, particularly, in the description of three- and four- dimensional
particle with Majorana spectrum \cite{maj},
$4d$ massless particles \cite{pl}.
The relation of the mentioned massless particle model
with $W-$ algebras has also been established \cite{rr}.
However, such systems were not studied completely
even in the four- dimensional space-time,
where only the first and second extrinsic curvatures were
considered. \\
In this note we attempt to fill this gap and
consider the simplest four-dimensional analog of (\ref{polac})
given by the action \cite{pc}
\begin{equation}
{\cal S} =\int(c_0+cK_3)d{\tilde s},\quad
d{\tilde s}=|d{\bf x}|\equiv sd\tau\neq 0
\label{0}\end{equation}
where $\tau$ is an arbitrary evolution parameter
and $K_3$ is the torsion (highest curvature) of a worldline in $4d$ space.
We will show that this system possesses interesting properties
which make it drastically different from other
four- and three- dimensional massive particle systems depending
on extrinsic curvatures:
\begin{itemize}
\item it has a zero spin;
\item it possesses, in addition to reparametrization invariance, the
extra gauge degrees of freedom: its classical trajectories are
restricted by the condition
$$
\frac{K^2_2K_3}{K^2_1}=|\alpha|, \quad \alpha={c_0}/{c}
$$
while the mass spectrum is described
the conformal
mechanics with the energy ${\cal E}$,
$$ dq\wedge dp, \quad {\cal H}=\frac{p^2}{2}\pm\frac{\alpha^2}{2q^2},
\quad
-2{\cal E}/\alpha^2= \left\{
\begin{array}{cc}
M^2/c^2_0\mp 1,&
{\rm if}\;\;\alpha < 0,\\
\pm M^2/c^2_0 \pm 1,&
{\rm if}\;\;\alpha >0
\end{array}\right. ,
$$
where
$q=K_1/K_2$, and $M$ denotes the mass of the system.\\
When $\alpha<0$, the upper sign corresponds to the
time-like trajectories, while other solutions corresponds
to the space-like ones.
\item When $\alpha<-1/2$, the solutions
with space-like trajectories possess a {\it discrete} spectrum with
massive and tachionic sectors,
while the solutions with time-like trajectories possess continuous
spectrum containing massive, massless, and tachionic solutions.
\end{itemize}
The paper is arranged as follows:
In {\it Section 2} we construct the Hamiltonian system corresponding
to the model under consideration (\ref{0}).
We show that the system possesses an extended gauge invariance and
a zero spin.
The geometry of its classical trajectories and quantum spectrum
(in the Euclidean space) are defined by the one-dimensional
conformal mechanics (with a repulsive potential).
In {\it Section 3} we reformulate
the Hamiltonian system constructed in Section 2 in the Minkowski space
and consider the properties of its mass spectrum.
\setcounter{equation}{0}
\section{Hamiltonian formulation}
In this section we give
the Hamiltonian formulation of the model (\ref{0}).
Recall
that the extrinsic curvatures $K_a$ of a (non-null) curve in
four-dimensional
space can be defined via the Frenet equations for the moving frame
${\bf e}_a$ :
\begin{eqnarray}
&{\dot{\bf x}}=s{\bf e}_1,\quad {\dot{\bf e}}_a=k_a^{\;b}{\bf e}_b,
\quad{\bf e}_a{\bf e}_b=\eta_{ab},\quad a,b,c=1,2,3,4; &\label{f}\\
& k_{a}^{\; c}\eta_{cb}=
\left(
\begin{array}{cccc}
0 &k_1 &0 &0\\
-k_1&0 &k_2 &0\\
0 &-k_2&0 &k_3\\
0 &0 &-k_3&0
\end{array}\right),\quad k_{a-1}=sK_{a-1},& \label{mat}
\end{eqnarray}
where ${\bf x}$ are coordinates of four-dimensional space,
and ${\bf e}_a$ are elements of the moving frame.\\
While $K_1, K_2$ are positive quantities, the sign of the
highest curvature $K_{3}$ (torsion) is not uniquely defined
by the Frenet equations \cite{postnicov}.
Without loss of generality we assume below that $K_3>0$.
In the Euclidean space $\eta_{ab}=\delta_{ab}$, so the
Frenet equations read
\begin{equation}
{\dot{\bf e}}_a=k_a{\bf e}_{a+1}-k_{a-1}{\bf e}_{a-1},\quad
{\bf e}_{0}={\bf e}_5\equiv 0.
\label{ffe}\end{equation}
One can transform the Frenet equations in the Euclidean space
into those in the Minkowski space,
performing
the following transition
\begin{equation}
({\bf e}_{\underline a},\; k_{\underline a},\;
k_{{\underline a}-1}, s)\to
(i{\bf e}_{\underline a},\; ik_{\underline a},\;
ik_{{\underline a}-1},(-i)^{\delta_{1{\underline a}}}s)
\label{tr}\end{equation}
for some index ${\underline a}$. \\
Indeed, this transformation preserves the form of the
matrix in (\ref{mat}),
while the element ${\bf e}_{\underline a}$ becomes time-like:
${\bf e}^2_{\underline a}=-1$.
So, we can give our basic derivations for the Euclidean case,
reformulating them for the Minkowski space
for a final analysis.
It follows from (\ref{ffe})
that
\begin{equation}
s=\sqrt{{\dot{\bf x}}^2},\quad
k_a={\dot{\bf e}}_a{\bf e}_{a+1}=\sqrt{{\dot{\bf e}_a}^2-k^2_{a-1}}.
\label{cf}\end{equation}
Thus, the Lagrangian appearing in the action (\ref{0})
in the {\it Euclidean} space
can be replaced by the following
one
\begin{equation}
L=c_0s+c\sqrt{{\dot{\bf e}}_3^2-k^2_2}+{\bf p}({\dot{\bf x}}-s{\bf e}_1)+
\sum_{i}{\bf p}_{i-1}({\dot{\bf e}}_{i-1}-k_{i-1}{\bf e}_{i}+
k_{i-2}{\bf e}_{i-2})-\sum_{i,j}d_{ij}({\bf e}_i{\bf e}_j-\delta_{ij}),
\label{l}\end{equation}
where
${\bf x}, {\bf p},{\bf e}_i,{\bf p}_{i-1}, s, k_i, d_{ij}$
are independent variables,
$i,j=1,2,3$.\\
Now we can perform the Legendre transformation for
this Lagrangian
(referring for
details to \cite{tmp}).\\
The variables ${\bf p}_{i-1}$ represent the momenta
conjugated to
${\bf e}_{i-1}$,
whereas the momenta conjugated to $(s, k_{i-1}, d_{ij})$
lead to the trivial constraints
\begin{equation}
p^s\approx 0,\quad p^{i-1}\approx 0,\quad p^{ij}\approx 0.
\label{erunda}\end{equation}
Setting $k_3\neq 0$ we find that the momentum
conjugated to ${\bf e}_3$ is of the form
\begin{equation}
{\bf p}_3=c\left({\dot{\bf e}}^2_3-k^2_{2}\right)^{-1/2}
{\bf{\dot e}}_3.
\label{pN}\end{equation}
Taking into account (\ref{cf}), we get the constraints
\begin{equation}
{\bf p}_3{\bf e}_{3}\approx 0,\quad
{\bf p}_3{\bf e}_{1}\approx 0,
\quad{\bf p}^2_3- ({\bf p}_3{\bf e}_{2})^2-c^2\approx 0.
\label{PhiNN0}\end{equation}
Then the construction of primary
Hamiltonian system becomes straightforward.
To simplify the resulting system, one can stabilize
trivial primary constraints
(\ref{erunda}) and exclude
them from our considerations, which makes
the variables $s, k_{i-1}, d_{ij}$ lagrangian multipliers.
Without
loss of generality one can
also impose the gauge conditions (see for a details \cite{tmp})
$${\bf p}_3{\bf e}_2\approx 0,\quad
{\bf p}_{2}{\bf e}_{2}\approx 0, \quad
{\bf p}_{2}{\bf e}_{1}\approx 0. $$
After these manipulations we get
the Hamiltonian system
\begin{equation}
\begin{array}{c}
\omega=d{\bf p}\wedge d{\bf x}+
\sum_{i} d{\bf p}_i\wedge d{\bf e}_i, \\
\;\;\\
{\cal H}=s({\bf p}{\bf e}_1 -c_0)+\sum_{i} k_{i-1}\phi_{i-1.i}+
{k_3\over 2c}({\bf p}^2_3-c^2)+\sum_{i,j}d_{ij}
({\bf e}_i{\bf e}_j-\delta_{ij}),
\end{array}
\label{ss}
\end{equation}
with primary constraints
\begin{eqnarray}
&{\bf p}_i{\bf e}_{j}\approx 0,\quad
i\geq j &\label{primary}\\
&{\bf e}_i{\bf e}_j-\delta_{ij}\approx 0, &
\label{u}\\
&{\bf p}{\bf e}_1 -c_0\approx 0,&\label{phi0}\\
&\phi_{i-1.i}\equiv{\bf p}_{i-1}{\bf e}_{i}-
{\bf p}_{i}{\bf e}_{i-1} \approx 0,
\quad{\bf p}^2_3-c^2\approx 0,&\label{phi}
\end{eqnarray}
where variables $s,k_i, d_{ij}$ play the role of
Lagrangian multipliers.\\
Let us stabilize the primary constraints (\ref{primary})-(\ref{phi}).\\
Stabilization of (\ref{primary}) gives the following fixation
of the Lagrangian multipliers $d_{ij}$:
\begin{equation}
2d_{i.j}=k_3c\delta_{i.3}\delta_{j.3}-sc_0\delta_{1.i}\delta_{1.j},
\label{d}\end{equation}
so that the equations of motion read
\begin{equation}
\begin{array}{c}
\dot{\bf x}={\bf e}_1,\\
{\dot{\bf e}}_1=k_1{\bf e}_2,\\
{\dot{\bf e}}_2=-k_1{\bf e}_1+k_2{\bf e}_3,\\
{\dot{\bf e}}_3=-k_2{\bf e}_2+k_3{\bf p}_3/c\\
{\dot{\bf p}}_3=-k_3c{\bf e}_3-k_2{\bf p}_2,\\
{\dot{\bf p}}_2=-k_1{\bf p}_1+k_2{\bf p}_3,\\
{\dot{\bf p}}_1=-s{\bf p}+k_1{\bf p}_2+sc_0{\bf e}_1, \\
{\dot{\bf p}}=0.
\end{array}\label{hem}\end{equation}
Stabilizing the remaining primary constraints,
we get the following first-stage secondary constraints
\begin{equation}
{\bf p}{\bf e}_2\approx 0,
\quad{\bf p}_{1}{\bf e}_{3}\approx 0,
\quad {\bf p}_3{\bf p}_2\approx 0.
\label{secondary}\end{equation}
From the ortogonality of (${\bf e}_i, {\bf p}_3/c$) to ${\bf p}_2$,
which follows from (\ref{phi}),(\ref{primary}),(\ref{secondary})
we conclude that
\begin{equation}
{\bf p}_2\approx 0.
\label{secondary1}\end{equation}
Stabilizing the first and second constraints from (\ref{secondary})
and the constraint (\ref{secondary1}), we get
\begin{eqnarray}
&k_2=qk_1, \quad k_3=sc_0/cq^2, &\label{multiplyers}\\
& {\bf p}_1=q{\bf p}_3,&\label{tertiary}
\end{eqnarray}
where
\begin{equation}
1/q\equiv{{\bf p}{\bf e}_3\over c_0}\neq 0.
\end{equation}
Then we get
\begin{equation}
{\bf p}=c_0{\bf e}_1+c_0{\bf e}_3/q+
p{\bf p}_3,\label{p}
\end{equation}
where
\begin{equation}
p\equiv {\bf p}{\bf p}_3/c^2.\label{pi}
\end{equation}
Consistency of the equations of motion (\ref{hem}) with the Frenet equations
implies that
\begin{equation}
{\bf e}_a=\{{\bf e}_i,\;{\bf e}_4\equiv{\bf p}_3/c\}:
\quad {\bf e}_a{\bf e}_b= \delta_{ab}.\label{on}\end{equation}
Therefore, the initial Hamiltonian system can be formulated purely in
terms of the moving frame ${\bf e}_a$, coordinates ${\bf x}$, and momentum
${\bf p}$, while the relations (\ref{on}), (\ref{p})
play the role of constraints.
Taking into account the constraint (\ref{p})),
we get the
expressions for the rotation generators:
\begin{equation}
{\bf J}={\bf p}\vee{\bf x}+\sum_{i}{\bf p}_i\vee{\bf e}_i=
{\bf p}\vee ({\bf x}-q{\bf p}_3/c_0)\label{J}\end{equation}
and the Casimirs
\begin{equation}
{\bf p}^2=(cp)^2+c^2_0(1+1/q^2),\quad
{\bf W}^2\equiv ({\bf p}\vee {\bf J})^2=0.
\end{equation}
Thus, {\it the system possesses a zero spin
despite the dependence of the initial Lagrangian
on higher derivatives.}\\
The expression for rotation generators (\ref{J}) hints us to introduce the
``effective" coordinate
\begin{equation}{\bf X}\equiv {\bf x}-q{\bf p}_3/c_0\;:
\quad {\bf{\dot X}}=s{\bf p}/c_0,\quad {\bf p}=const
\label{X}\end{equation}
whose evolution equations look similar to the convenient relation
between velocity and momentum of a massive particle
(note, that in spite of these relations
the mass of the system is not equal to $c_0$).
The reduction of the initial Hamiltonian system (\ref{ss})
by the constraints
(\ref{on}), (\ref{p}), (\ref{tertiary}) leads to
the following unconstrained system
\begin{equation}
\omega^{red}= d{\bf p}\wedge d{\bf X}+\frac{c^2}{c_0}
{dp\wedge dq},\quad
{\cal H}^{red}=\frac{sc^2}{c_0}\left(
p^2/2+\frac{c^2_0}{2c^2q^2}-\frac{{\bf p}^2- c^2_0}{2c^2}\right).
\label{fd}\end{equation}
The external curvatures of the system are related with the modular
``coordinate" $q$ as follows
\begin{equation}
\frac{K_2}{K_1}=q, \quad K_3=\frac{c_0}{cq^2},\quad\Rightarrow\quad
\frac{K^2_2K_3}{K^2_1}=\frac{c_0}{c}.
\label{Kq}\end{equation}
So, {\it the system under consideration
possesses extended gauge invariance
since the trajectories
with the same ratio $K_{1}/K_{2}$ are gauge equivalent.}\\
Reducing the system (\ref{fd}) by ${\bf p}$, to exclude
the trivial part of dynamics (\ref{X}),
we get that the
evolution of the parameter $q$ is described by the textbook
mechanical system \cite{ll}
\begin{equation}
dp\wedge dq,\quad
{\cal H}_0=p^2/2+\frac{c^2}{2c^2_0q^2},\quad {\cal E}=
\frac{{\bf p}^2-c^2_0}{2c^2}.
\label{cm}\end{equation}
This system can be immediately integrated at the classical
level
$$ q^2=\frac{({\bf p}^2-c^2_0){\tilde s}^2}{2c^2}+
\frac{2}{{\bf p}^2-c^2_0},\quad{\bf p}^2>c^2_0,$$
as well as at the quantum one.
Thus, adding of the torsion term increases the absolute value of momenta
of the system with respect to the torsionless system,
${\bf p}^2\geq c^2_0$.
The system (\ref{cm}) is known in literature as a conformal mechanics,
due to the symmetry of its action under conformal transformations
generated by
\begin{equation}
{\cal H}_0=p^2/2+\frac{c^2}{2c^2_0q^2},
\quad{\cal D}=p q,\quad {\cal K}={q^2}/{2}\;.
\end{equation}
The ``energy" spectrum is continuous and has the lowest bound ${\cal E}=0$
which is not normalizable.
In \cite{nc} the conformal symmetry of this system
has been used for solving the problem of the ground state.
Recently this mechanism has been found to be adequate
to the problem of motion of charged particle in the field
of a charged black hole \cite{k}. Due to this
observation the interest to the conformal mechanics
and its supergeneralizations \cite{pashnev},
is renewed in the context of
study of probe $D0$-brane dynamics in the external $D$-brane field
(see e.g.\cite{branes} and refs therein).
Conformal mechanics arises in our model a different context:
it defines local gauge symmetry of the particle system.
\setcounter{equation}{0}
\section{Transition to Minkowski space}
In the previous Section we constructed the constrained Hamiltonian system
corresponding to the action (\ref{0}) in the Euclidean space.
We found that the
evolution and the geometry of this system are described in terms of the
non-relativistic conformal mechanics with repulsive potential.
The purpose of this Section is to consider the relativistic aspects of
this system, i.e. to investigate the model (\ref{0}) in the
Minkowski space.
As we have mentioned in the beginning of Section 2,
we can
use the results concerning the Euclidean case,
since the Frenet formulae in the Euclidean space can be transformed
into the ones in
Minkowski space with the use of transition (\ref{tr}),
where ${\underline a}$ denotes
an element of moving frame which becomes time-like, i.e.
${\bf e}^2_{\underline a}=-1$.
Correspondingly,
for the transition to {\it time-like} trajectories we have to choose
${\underline a}=1$, while for the transition to
{\it space-like} trajectories
we have to choose ${\underline a}\neq 1$.
We do not consider here systems with {\it light-like} trajectories.
For convenience we will use the notation
\begin{equation}
{\bf p}^2\equiv-M^2,\quad \alpha=\frac{c_0}{c},
\end{equation}
so that $M^2>,=,< 0$ corresponds to
the massive, massless, and tachionic sectors of the system, respectively.
To reformulate the system under consideration in
the Minkowski space we have to supply the transition
(\ref{tr}) with appropriate
transformations of characteristic constants $c_0, c$ and momenta
${\bf p}_i$ which preserve the form of
the initial action (\ref{0})
and the initial Hamiltonian system, namely,
\begin{equation}
\begin{array}{cccc}
{\underline a}=1:&({\bf e}_1,{\bf p}_1, k_1, s, c_0)&\to &
( i{\bf e}_1, -i{\bf p}_1, ik_1,-is, ic_0);\\
{\underline a}=2:&({\bf e}_2,{\bf p}_2, k_2, k_1)&\to &
( i{\bf e}_2, -i{\bf p}_2, ik_2, ik_1),\\
{\underline a}=3:&
({\bf e}_3,{\bf p}_3, k_3, k_2, c)&\to &
( i{\bf e}_3, -i{\bf p}_3, ik_3, ik_2, -ic),\\
{\underline a}=4:&
(c, k_3)&\to &( -ic, ik_3).
\end{array}
\label{htr}\end{equation}
Indeed, it is easy to see that it is only the
ortonormality condition
${\bf e}_a{\bf e}_b=(-1)^{\delta_{{\underline a}a}}\delta_{ab}$
that is changes in (\ref{ss}) under this transformation.
Consequently, the reduced system (\ref{cm}) with appropriately
changed parameters describes the effective Hamiltonian system corresponding
to the action (\ref{0}) in the Minkowski space.
The transition (\ref{htr}) induces the following
transformation of the parameters and coordinates
of the reduced system (\ref{cm})
\begin{equation}
(p,q,\alpha, c_0, s)\to
\left\{
\begin{array}{cc}
(p, iq, i\alpha, ic_0, -is)&{\rm if } \;{\underline a}=1\\
(p,q,\alpha, c_0, s)& {\rm if } \;{\underline a}=2\\
(ip, -iq, i\alpha, c_0, s ) &{\rm if } \;{\underline a}=3\\
(-p, q, i\alpha, c_0, s ) &{\rm if } \;{\underline a}=4.
\end{array}\right.
\end{equation}
Thus, the reduced system reads
\begin{equation}
dp\wedge dq,\quad
\epsilon_{{\underline a}}s\left(
\frac{p^2}{2} +
\frac{\epsilon_{{\underline a}}\alpha^2}{2q^2}
-{\cal E}_{\underline a}\right)\approx 0,
\end{equation}
where
$$
{\cal E}_{{\underline a}}=
-(-1)^{\delta_{4{\underline a}}}\frac{\alpha^2}{2}
({M^2}/{c^2_0}+(-1)^{\delta_{1{\underline a}}}),\quad
\epsilon_{{\underline a}}=\left\{
\begin{array}{cc}
1,& {\rm if }\;\;{\underline a}=1,2\\
-1,& {\rm if }\;\;{\underline a}=3,4.
\end{array}\right.
$$
The
expressions for curvatures (\ref{Kq}) read
\begin{equation}
\frac{K_2}{K_1}=(-1)^{\underline a}q, \quad
K_3=(-1)^{\underline a}\frac{\alpha}{q^2},\quad\Rightarrow\quad
\frac{K^2_2K_3}{K^2_1}=(-1)^{\underline a}\alpha .
\label{Kqm}\end{equation}
Due to positivity of curvatures $K_i$, we conclude that
\begin{itemize}
\item ${\underline a}=1,3$, $\quad q<0$,\hspace{0.5cm} if\hspace{0.5cm}
$\alpha<0$,
\item
${\underline a}=2,4$, $\quad 0<q$,\hspace{0.5cm} if\hspace{0.5cm} $\alpha>0$.
\end{itemize}
Thus, in both cases the system possesses the solutions described
by conformal mechanics with a repulsive (${\underline a}=1,2$) and an
attractive (${\underline a}=3,4$) potentials.
The classical solutions of these systems are of the
form
$$
q^2=\left\{
\begin{array}{cc}
{\cal E}_{\underline a}{\tilde s}^2+(-1)^{[{\underline a}]}
{\alpha^2}/{{\cal E}_{\underline a}},&{\rm if} \;\; {\cal E}\neq 0,\\
2\alpha {\tilde s},&{\rm if}\;\; {\cal E}=0,\;\;
{\underline a}=3,4
\end{array}\right.
$$
When the potential
is attractive (${\underline a}=3,4$),
the parameter ${\tilde s}$ is defined on the domain
$$
{\tilde s}\in\left\{
\begin{array}{cc}
] -|\alpha/{\cal E}|,\;\;|\alpha/{\cal E}|[, & {\rm if}\; {\cal E}<0,\\
]|\alpha/{\cal E}| ,\;\;\infty [,& {\rm if} \;{\cal E}>0.
\end{array}\right.
$$
While the quantum spectrum of the mechanics with a repulsive potential
is continuous, the spectrum of conformal mechanics with an attractive
potential
can be both continuous and discrete.
The discrete spectrum corresponds to the
strongly attractive potential ($|\alpha|>1/2$)
and has an infinite number of energy levels \cite{pr}.\\
We first consider the systems with repulsive potential
\begin{equation}
dp\wedge dq,\quad
{\cal H}_0=\frac{p^2}{2} +
\frac{\alpha^2}{2q^2},\quad \frac{2{\cal E}}{\alpha^2}=\left\{
\begin{array}{cc}
(1-{M^2}/{c^2_0}),& {\rm if}\;\; \alpha<0,\;\;{\underline a}=0\\
-(1+{M^2}/{c^2_0}),&{\rm if}\;\; \alpha>0,\;\;{\underline a}=1
\end{array}\right.
\end{equation}
Notice that while in the first case trajectories are {\it time-like},
those are {\it space-like} in the second case. \\
Since the energy ${\cal E}$ is positive both in classical and quantum cases,
we get the restrictions on
the admissible value of the mass
\begin{itemize}
\item $M^2<c^2_0$,$\;$ for $\alpha<0$ and time-like trajectories
(${\underline a}=1$),
\item $M^2<-c^2_0$,$\;$ for $\alpha>0$, ${\underline a}=2$.
\end{itemize}
Thus, in the first case there are massive, massless, and tachionic
states, while in the second case the system is tachionic.\\
Now we consider the systems with an attractive potential. In this case
all trajectories are space-like and are defined by the mechanics
\begin{equation}
dp\wedge dq,\quad
{\cal H}_0=\frac{p^2}{2} -
\frac{\alpha^2}{2q^2},\quad \frac{2{\cal E}}{\alpha^2}=
\left\{
\begin{array}{cc}
-(1+{M^2}/{c^2_0})&
{\rm if}\;\; \alpha<0,\; {\underline a}=3\\
(1+{M^2}/{c^2_0})&
{\rm if} \;\;\alpha>0,\; {\underline a}=4
\end{array}\right.
\end{equation}
The spectrum of conformal mechanics with an attractive potential is
continuous for ${\cal E}>0$ and for ${\cal E}\leq 0, \alpha^2<1/4$
(see \cite{ll}).
If the potential is ``strongly attractive" $\alpha^2>1/4$,
and ${\cal E}<0$,
the system has a dicrete spectrum with an
infinite number of bound states given
by the expression \cite{pr}
$$
{\cal E}_n=-\hbar^2 B^2\exp{\frac{-2\pi n}{\sqrt{\alpha^2-1/4}}},\quad
n=0,\pm1,\pm2,\ldots,
$$
where $B$ is an undefined "phase" factor. \\
So,
\begin{equation}
M^2_n/c^2_0=-1+
(-1)^{{\rm sgn}\alpha}\left(\frac{B\hbar}{\alpha}\right)^2
\exp{\frac{-2\pi n}{\sqrt{\alpha^2-1/4}}},\quad |\alpha|>1/2.
\end{equation}
Therefore, for $\alpha>0$ the discrete branch corresponds to pure
tachionic states, while for $\alpha<0$ it contains massive, massless,
and tachionic sectors. \\
{\large Acknowledgments.} I am thankful to M.Plyushchay,
who suggested to investigate this model and gave numerous advices
and C.Sochichiu for valuable discussions and useful comments.
This work has been partially supported by
grants INTAS-RFBR No.95-0829, INTAS-96-538 and INTAS-93-127-ext.
|
\section{Introduction}
\noindent
Semileptonic 4-body decays of $B$-mesons with emission of a single pion
have been studied in detail by many authors \cite{BL,korner,bl4}.
Recently we investigated the possibility of probing direct CP violation in
the decay $B^\pm\to D\pi l^\pm \nu$ \cite{bl4}
in extensions of the Standard Model (SM), where we extended the weak
charged current by including a scalar-exchange interaction
with a complex coupling, and considered as specific models the
multi-Higgs doublet (MHD) model and the scalar-leptoquark (SLQ) models.
In the present work, we investigate the same possibility in the decay of
$B^\pm\to (\pi^+\pi^-) l^\pm \nu$. In this case we find there may be
direct CP violation even within the SM.
As is well known, in order to observe direct CP violation effects, there should
exist interferences not only through weak CP-violating phases but also with
different CP-conserving strong phases.
In the decay of $B^\pm\to \pi\pi l^\pm \nu$, we consider it as a two-stage process:
$B\to (\sum_{i} M_i\to \pi\pi)l\nu$, where $M_i$ stands for an intermediate
state which is decaying to $\pi^+ + \pi^-$.
In this picture the CP-conserving phases may come from the absorptive parts
of the intermediate resonances.
Here we try to include as many as possible intermediate states decaying
to $\pi^+ + \pi^-$,
so that they could represent a pseudo complete set of the relevant decay.
The candidates in $b\to u$ transition are $\rho$, $f_0$ and $f_2$ mesons,
which decay dominantly to $2\pi$ mode (See Table.~1).
Furthermore, we find that even in $b\to c$ transition
a $D^0$ meson can decay to $\pi^++\pi^-$,
although its branching fraction is very small compared to those of
$u\bar{u}$ states.
However, we can find that the contribution through an intermediate $D$
meson is not negligible because of CKM favored nature of $b\to c$ transition
compared to the $b\to u$ one of $u\bar{u}$ states, $\rho$, $f_0$ and $f_2$
mesons.
If we include $D$ meson as an intermediate state as well as the $u\bar{u}$
states, direct CP violation may arise even within the SM through their relative
weak phases of the different CKM matrix elements ($V_{cb}$ and $V_{ub}$).
Therefore, we first consider CP violation within the SM by including
$\rho$, $f_0$, $f_2$ and $D$ mesons\footnote{
Here we are including fully known resonances only,
and neglecting possible non-resonant $2\pi$ decays.
A significant experimental enhancement can be made, if we use the reduced
kinematic region around
1.4 GeV $\leq \sqrt{(p_\pi+p_\pi)^2} \leq$ 1.9 GeV (See Table.~1).}
as intermediate states decaying to $\pi^+\pi^-$.
Next we also consider CP violations in extensions of the SM,
in which we use a cutoff to the final state $\pi\pi$ invariant mass
so that the effects of $D$ meson cannot enter, thus ensuring
that the result is solely from new physics.
In Section 2, we present our formalism dealing with $B\to \pi\pi l \nu$ decays
within the SM and in its extensions,
and the observable asymmetries are considered in Section 3.
Section 4 contains our numerical results and conclusions.
Presented in Appendix are all the relevant formulae we use here.
\begin{table}
{Table~1}. {Properties and branching ratios of $\pi^+\pi^-$ resonances}\par
\begin{tabular}{|c|l|c|c|c|c|}
\hline
Label $i$ & $\qquad\quad M_i$ & $J^P$ & $m_i$ (MeV) & $\Gamma_i$ (MeV) &
${\cal BR}_i(M_i\to \pi^+ \pi^-)$\\
\hline
$0$ & $M_0=f_0(980)$ & $0^+$ & $980$ & $40\sim 100$ & 0.52\\
$0^\prime$ & $M_{0^\prime}=f_0(1500)$ & $0^+$ & $1500$ & $112$ & 0.3\\
$1$ & $M_1=\rho (770)$ & $1^-$ & $770$ & $151$ & 1\\
$1^\prime$ & $M_{1^\prime}=\rho (1700)$ & $1^-$ & $1700$ & $240$ & 0.3\\
$2$ & $M_2=f_2 (1270)$ & $2^+$ & $1275$ & $186$ & 0.56\\
$3$ & $M_3=D^0$ & $0^-$ & $1865$ & $1.59\times 10^{-9}$ & $1.53\times 10^{-3}$\\
\hline
\end{tabular}
\end{table}
\section{Theoretical Details of Decay Amplitudes}
\begin{center}
{\bf A. Within the Standard Model}
\end{center}
\noindent
The decay amplitudes for the processes of Fig.~1, with $M_i$ listed in Table 1,
\begin{eqnarray}
B^-(p_B)\to M_i(p_i,\l_i)+W^*(q)\to
\pi^+(p_+)+\pi^-(p_-)+l^-(p_l,\l_l)+\bar{\nu}(p_\nu)
\end{eqnarray}
are expressed as
\begin{eqnarray}
{\cal A}^{\lambda_l}&=&-\frac{G_F}{\sqrt{2}}\sum_{i}\sum_{\lambda_i}
V_i c_i \langle l^-(p_l,\lambda_l)\bar{\nu}(p_\nu)|j^{\mu\dagger}|0\rangle
\langle M_i(p_i,\lambda_i)|J_{i\mu}|B^-(p_B)\rangle \nonumber\\
& &\times \Pi_i(s_{_M})\langle \pi^+(p_+)\pi^-(p_-)\|M_i(p_i,\lambda_i)\rangle ,
\end{eqnarray}
where $\l_i=0$ for spin $0$ states ($f_0$ and $D$),
$\l_i=\pm 1,0$ for spin $1$ states ($\rho$),
$\l_i=\pm 2,\pm 1,0$ for spin $2$ states ($f_2$),
and $\l_l$ is the lepton helicity, $\pm\frac{1}{2}$.
\begin{figure}[ht]
\hbox to\textwidth{\hss\epsfig{file=fig1.ps,height=9cm,width=8cm,angle=-90}\hss}
\caption{Diagrams for $B\to M_iW^*\to \pi^+\pi^- l^-\bar{\nu}_l$ decays.}
\label{fig:diagram}
\end{figure}
The leptonic current is
\begin{eqnarray}
j^\mu = \bar{\psi}_\nu \gamma^\mu(1-\gamma_5)\psi_l,
\end{eqnarray}
and for the hadronic currents we have
\begin{eqnarray}
J_i^\mu = \bar{\psi}_u \gamma^\mu(1-\gamma_5)\psi_b, \;\; V_i=V_{ub},
\;\; c_i=\frac{1}{\sqrt{2}}\;\;\;
{\rm for}\;\; {\rm label}~~i=0^{^{(\prime)}},1^{^{(\prime)}},2;
\end{eqnarray}
and
\begin{eqnarray}
J_3^\mu = \bar{\psi}_c \gamma^\mu(1-\gamma_5)\psi_b, \;\; V_3=V_{cb},
\;\; c_3=1\;\;\; {\rm for}\;\; {\rm label}~~i=3,
\end{eqnarray}
where $c_i$ stands for the isospin factor especially due to $u\bar{u}$-mesons.
We assume that the resonance contributions of the intermediate states
can be treated by the Breit-Wigner form, which is written
in the narrow width approximation as
\begin{eqnarray}
\Pi_i (s_{_M})=\frac{\sqrt{m_i\Gamma_i/\pi}}{s_{_M}-m_i^2+im_i\Gamma_i},
\end{eqnarray}
where $s_{_M}=(p_+ + p_-)^2$ and the $m_i$'s and $\Gamma_i$'s are the masses
and widths of the resonances respectively (See Table 1).
For the decay parts of the resonances we use \cite{cancel}
\begin{eqnarray}
\langle \pi^+(p_+)\pi^-(p_-)\|M_i(p_i,\lambda_i)\rangle
=\sqrt{{\cal BR}_i}Y^{\lambda_i}_{\lambda_i max} (\theta^*,\phi^*),
\end{eqnarray}
where $Y^m_l(\theta,\phi)$ are the $J=l$ spherical harmonics listed in Appendix B,
and the angles $\theta^*$
and $\phi^*$ are those of the final state $\pi^-$ specified in the $M_i$
rest frame (See Fig. 2c).
The couplings of $M_i$ to $\pi\pi$ are effectively taken into account by
the branching fractions, ${\cal BR}_i(M_i\to \pi^+\pi^-)$.
\begin{figure}[tb]
\hbox to\textwidth{\hss\epsfig{file=fig2.ps,height=15cm,width=5.5cm,angle=-90}\hss}
\caption{The decay $B\to M_iW^*\to (\pi^+\pi^-)(l\bar{\nu})$ viewed from the
(a) $B^-$, (b) $W^*$ and (c) $M_i$ rest frames.}
\label{fig:frame}
\end{figure}
In order to obtain the full helicity amplitude of the $B\to \pi\pi l\nu$ decay,
we first consider the amplitude of $B\to M_i l \bar{\nu}_l$ \cite{bl3},
denoted as ${\cal M}^{\l_l}_{\l_i}$:
\begin{eqnarray}
{\cal M}^{\l_l}_{\l_i}=-\frac{G_F}{\sqrt{2}}
V_i c_i \langle l^-(p_l,\lambda_l)\bar{\nu}(p_\nu)|j^{\mu\dagger}|0\rangle
\langle M_i(p_i,\lambda_i)|J_{i\mu}|B^-(p_B)\rangle .
\end{eqnarray}
We express the matrix elements
${\cal M}^{\l_l}_{\l_i}$ into the following form:
\begin{eqnarray}
{\cal M}^{\l_l}_{\l_i}=\frac{G_F}{\sqrt{2}}V_i c_i
\sum_{\l_W}\eta_{\l_W}L^{\l_l}_{\l_W}H^{\l_i}_{\l_W},
\label{smamp}
\end{eqnarray}
where for the decays $B \to M_i W^*$ and $W^* \to l \bar{\nu}$, respectively,
\begin{eqnarray}
H^{\l_i}_{\l_W}&=&\epsilon^*_{W\mu}\langle M_i(p_i,\l_i)|J^\mu_i|B^-(p_B)\rangle ,\nonumber\\
L^{\l_l}_{\l_W}&=&\epsilon_{W\mu}\langle l^-(p_l,\l_l)\bar{\nu}(p_\nu)|j^{\mu\dagger}|0\rangle ,
\end{eqnarray}
in terms of the polarization vectors $\epsilon_W\equiv \epsilon(q,\l_W)$ of the virtual $W$.
These $\epsilon_W$'s satisfy the relation
\begin{eqnarray}
-g^{\mu\nu}=\sum_{\l_W}\eta_{\l_W}\epsilon^\mu_W \epsilon^{*\nu}_W,
\label{metric}
\end{eqnarray}
where the summation is over the helicities $\l_W =\pm 1,0,s$ of the virtual $W$,
with the metric $\eta_\pm=\eta_0=-\eta_s=1$.
We evaluate the leptonic amplitude $L^{\l_l}_{\l_W}$ in the rest frame of
the virtual $W$ (See Fig. 2b) with the $z$-axis chosen along the $M_i$
direction, and the $x$--$z$ plane chosen as the virtual $W$ decay plane,
with $(p_l)_x>0$.
Using the 2-component spinor technique \cite{hagiwara} and
polarization vectors given in Appendix B, we find
\begin{eqnarray}
&&L^-_\pm=2\sqrt{q^2}vd_\pm,\; L^-_0=-2\sqrt{q^2}vd_0,\; \hspace{1.1cm} L^-_s=0,\nonumber\\
&&L^+_\pm=\pm 2m_lvd_0,\; L^+_0=\sqrt{2}m_lv(d_+-d_-),\; L^+_s=-2m_lv,
\label{Lpm}
\end{eqnarray}
where
\begin{eqnarray}
v=\sqrt{1-\frac{m_l^2}{q^2}},\; d_\pm=\frac{1\pm\cos\theta_l}{\sqrt{2}},\;
~~{\rm and}~~d_0=\sin\theta_l.
\end{eqnarray}
Here we show only the sign of $\l_l$ as a superscript on $L$.
Note that the $L^+$ amplitudes are proportional
to the lepton mass $m_l$, and the scalar amplitude $L^-_s$ vanishes due to
angular momentum conservation.
For $B\to M_i$ transitions through the weak charged current
\begin{eqnarray}
J_i^\mu=V^\mu_i-A^\mu_i,
\end{eqnarray}
the most general forms of matrix elements are
\begin{eqnarray}
{\rm for}&f_0(0^+)&{\rm states}:\nonumber\\
&&\langle f_0(p_i)|V_\mu|B(p_B)\rangle =0,\nonumber\\
&&\langle f_0(p_i)|A_\mu|B(p_B)\rangle
=u_+(q^2)(p_B+p_i)_\mu+u_-(q^2)(p_B-p_i)_\mu;\nonumber\\
{\rm for}&\rho(1^-)&{\rm states}:\nonumber\\
&&\langle \rho(p_i,\epsilon_1)|V_\mu|B(p_B)\rangle
=ig(q^2)\epsilon_{\mu\nu\rho\sigma}\epsilon_1^{*\nu}(p_B+p_i)^\rho
(p_B-p_i)^\sigma,\nonumber\\
&&\langle \rho(p_i,\epsilon_1)|A_\mu|B(p_B)\rangle
=f(q^2)\epsilon^*_{1\mu}+a_+(q^2)(\epsilon_1^*\cdot p_B)(p_B+p_i)_\mu\nonumber\\
&&\hspace{4.5cm} +a_-(q^2)(\epsilon_1^*\cdot p_B)(p_B-p_i)_\mu;\nonumber\\
{\rm for}&f_2(2^+)&{\rm states}:\nonumber\\
&&\langle f_2(p_i,\epsilon_2)|V_\mu|B(p_B)\rangle
=ih(q^2)\epsilon_{\mu\nu\l\rho}\epsilon_2^{*\nu\alpha}p_{B\alpha}
(p_B+p_i)^\l(p_B-p_i)^\rho,\nonumber\\
&&\langle f_2(p_i,\epsilon_2)|A_\mu|B(p_B)\rangle
=k(q^2)\epsilon^*_{2\mu\nu}p_B^\nu +b_+(q^2)(\epsilon^*_{2\alpha\beta}
p_B^\alpha p_B^\beta)(p_B+p_i)_\mu\nonumber\\
&&\hspace{4.5cm} +b_-(q^2)(\epsilon^*_{2\alpha\beta}p_B^\alpha p_B^\beta)(p_B-p_i)_\mu;\nonumber\\
{\rm for}&D(0^-)&{\rm states}:\nonumber\\
&&\langle D(p_i)|V_\mu|B(p_B)\rangle =f_+(q^2)(p_B+p_i)_\mu+f_-(q^2)(p_B-p_i)_\mu,\nonumber\\
&&\langle D(p_i)|A_\mu|B(p_B)\rangle =0,
\label{BMamp}
\end{eqnarray}
where $\epsilon_1$ and $\epsilon_2$ are the polarization vectors of the spin 1
and spin 2 states, respectively.
Using the above expressions and the polarization vectors given in Appendix B,
we find non-zero $B\to M_iW^*$ amplitudes are
\begin{eqnarray}
{\rm for}&i=0,&\;\; H^0_{\l_W}\equiv S^0_{\l_W},\nonumber\\
&&S^0_0=-u_+(q^2)\frac{\sqrt{Q_+Q_-}}{\sqrt{q^2}},\;
S^0_s=-\left(u_+(q^2)\frac{(m_B^2-s_{_M})}{\sqrt{q^2}}+u_-(q^2)\sqrt{q^2}\right),
\end{eqnarray}
\begin{eqnarray}
{\rm for}&i=1^{^{(\prime)}},&H^{\l_1}_{\l_W}\equiv V^{\l_1}_{\l_W},\nonumber\\
&&V^0_0=-\frac{1}{2\sqrt{s_{_M}q^2}}
\left[f(q^2)(m_B^2-s_{_M}-q^2)+a_+(q^2)Q_+Q_-\right],\nonumber\\
&&V^{\pm 1}_{\pm 1}=f(q^2)\mp g(q^2)\sqrt{Q_+Q_-},\nonumber\\
&&V^0_s=-\frac{\sqrt{Q_+Q_-}}
{2\sqrt{s_{_M}q^2}}\left[f(q^2)+a_+(q^2)(m_B^2-s_{_M})+a_-(q^2)q^2\right],
\end{eqnarray}
\begin{eqnarray}
{\rm for}&i=2,&H^{\l_2}_{\l_W}\equiv T^{\l_2}_{\l_W},\nonumber\\
&&T^0_0=-\frac{1}{2\sqrt{6}}\frac{\sqrt{Q_+Q_-}}{s_{_M}\sqrt{q^2}}
\left[k(q^2)(m_B^2-s_{_M}-q^2)+b_+(q^2)Q_+Q_-\right],\nonumber\\
&&T^{\pm 1}_{\pm 1}=\frac{1}{2\sqrt{2}}\sqrt{\frac{Q_+Q_-}{s_{_M}}}
[k(q^2)\mp h(q^2)\sqrt{Q_+Q_-}],\nonumber\\
&&T^0_s=-\frac{1}{2\sqrt{6}}\frac{Q_+Q_-}{s_{_M}\sqrt{q^2}}
\left[k(q^2)+b_+(q^2)(m_B^2-s_{_M})+b_-(q^2)q^2\right],
\end{eqnarray}
\begin{eqnarray}
{\rm for}&i=3,&H^0_{\l_W}\equiv P^0_{\l_W},\nonumber\\
&&P^0_0=f_+(q^2)\frac{\sqrt{Q_+Q_-}}{\sqrt{q^2}},\;
P^0_s=f_+(q^2)\frac{(m_B^2-s_{_M})}{\sqrt{q^2}}+f_-(q^2)\sqrt{q^2},
\end{eqnarray}
where
\begin{eqnarray}
Q_\pm=(m_B\pm\sqrt{s_{_M}})^2-q^2.
\label{Qpm}
\end{eqnarray}
Here
\begin{eqnarray}
Q_+Q_-=\l(m_B^2,s_{_M},q^2)
\end{eqnarray}
gives the triangle function $\l(a,b,c)=a^2+b^2+c^2-2(ab+bc+ca)$.
Combining all the formulae, we can write the full helicity amplitudes of
$B^-\to \pi^+\pi^- l^-\bar{\nu}$ decays as
\begin{eqnarray}
{\cal A}^{\lambda_l}&=&V_{ub}\frac{G_F}{\sqrt{2}}\bigg[
\sum_{\l=0,s} \eta_\l L^{\l_l}_\l (\Pi_{f_0} S^0_\l Y^0_0 +
\xi \Pi_D P^0_\l \tilde{Y}^0_0+\Pi_\rho V^0_\l Y^0_1
+\Pi_{f_2} T^0_\l Y^0_2)\nonumber\\
&&+\sum_{\l=\pm 1} L^{\l_l}_\l (\Pi_\rho V^\l_\l Y^\l_1
+\Pi_{f_2} T^\l_\l Y^\l_2)\bigg],
\label{amp}
\end{eqnarray}
where
\begin{eqnarray}
\Pi_{f_0}&\equiv&\frac{1}{\sqrt{2}}(\sqrt{{\cal BR}_0}\Pi_0
+\sqrt{{\cal BR}_{0^\prime}}\Pi_{0^\prime})\nonumber\\
\Pi_{\rho}&\equiv&\frac{1}{\sqrt{2}}(\sqrt{{\cal BR}_1}\Pi_1
+\sqrt{{\cal BR}_{1^\prime}}\Pi_{1^\prime})\nonumber\\
\Pi_{f_2}&\equiv&\frac{1}{\sqrt{2}}\sqrt{{\cal BR}_2}\Pi_2\nonumber\\
\Pi_D&\equiv&\sqrt{{\cal BR}_3}\Pi_3,
\end{eqnarray}
and
\begin{eqnarray}
\xi=\frac{V_{cb}}{V_{ub}}.
\label{xi}
\end{eqnarray}
Note that we use $\tilde{Y}^0_0$ for the pseudo scalar
meson $D$, which is actually the same quantity as $Y^0_0=1/\sqrt{4\pi}$
for the scalar meson $f_0$
except that it changes sign under the parity transformation.
Concerning the parametrization of $\xi$,
other CKM factors, such as $V_{cd}^*$ from $D^0\to \pi^+\pi^-$ decay,
are already included in its branching fraction calculation.
And because we use implicitly Wolfenstein parametrization \cite{wolfenstein}
for CKM matrix, in which the complex phases are approximately in the elements
$V_{td}$ and $V_{ub}$ only,
the imaginary part of $\xi$ here comes only from the element $V_{ub}$.
The differential partial width of interest can be expressed as
\begin{eqnarray}
d\Gamma(B^- \to \pi^+ \pi^- l^- \bar{\nu}_l)
=\frac{1}{2m_B}\sum_{\l_l}|{\cal A}^{\l_l}|^2
\frac{(q^2-m_l^2)\sqrt{Q_+Q_-}}{256\pi^3m_B^2q^2}d\Phi_4,
\end{eqnarray}
where the 4 body phase space $d \Phi_4$ is
\begin{eqnarray}
d \Phi_4 \equiv ds_{_M} \cdot dq^2 \cdot d\cos\theta^* \cdot
d\cos\theta_l \cdot d\phi^*.
\end{eqnarray}
Kinematically allowed regions of the variables are
\begin{eqnarray}
&&4m_\pi^2\;<\;s_{_M}\;<\;(m_B-m_l)^2,\nonumber\\
&&m_l^2\;<\;q^2\;<\;(m_B-\sqrt{s_{_M}})^2,\nonumber\\
&&-1\;<\;\cos\theta^*,\;\cos\theta_l\;<\;1,\nonumber\\
&&0\;<\;\phi^*\;<\;2\pi .
\end{eqnarray}
Since the initial $B^-$ system is not CP self-conjugate, any genuine
CP-odd observable can be constructed only by considering both the $B^-$
decay and its charge-conjugated $B^+$ decay, and by identifying the CP
relations of their kinematic distributions.
Before constructing possible CP-odd asymmetries explicitly, we calculate
the decay amplitudes for the charge-conjugated process
$B^+\to \pi^+ \pi^- l^+ \nu_l$.
For the charge-conjugated $B^+$ decays,
the amplitudes can be written as
\begin{eqnarray}
{\bar{\cal A}}^{\l_l}&=&-\frac{G_F}{\sqrt{2}}\sum_{i}\sum_{\l_i}
V_i^* c_i \langle l^+(p_l,\lambda_l)\nu(p_\nu)|j^{\mu}|0\rangle
\langle \overline{M}_i(p_i,\lambda_i)|J^\dagger_{i\mu}|B^+(p_B)\rangle \nonumber\\
& &\times \Pi_i(s_{_M})\langle \pi^+(p_+)\pi^-(p_-)\|\overline{M}_i(p_i,\lambda_i)\rangle .
\end{eqnarray}
The leptonic amplitudes $\bar{L}^{\l_l}_{\l_W}$ are
\begin{eqnarray}
&&\bar{L}^+_\pm=-2\sqrt{q^2}vd_\mp,\; \bar{L}^+_0
=-2\sqrt{q^2}vd_0,\; \hspace{1.1cm} \bar{L}^+_s=0,\nonumber\\
&&\bar{L}^-_\pm=\pm 2m_lvd_0,\;\;\;\; \bar{L}^-_0=\sqrt{2}m_lv(d_+-d_-),\; \bar{L}^-_s=-2m_lv.
\end{eqnarray}
And the transition amplitudes $\overline{H}^{\l_i}_{\l_W}$ for $B^+\to \overline{M}_iW^*$
are given by a simple modification of the
amplitudes $H^{\l_i}_{\l_W}$ of the $B^-$ decays:
\begin{eqnarray}
\overline{H}^{\l_i}_{\l_W}=H^{\l_i}_{\l_W}\{g\to -g,\;h\to -h,\;f_\pm\to-f_\pm\}.
\end{eqnarray}
Then, we find the full amplitude for $B^+\to \pi^+\pi^-l^+\nu$:
\begin{eqnarray}
{\bar{\cal A}}^{\lambda_l}&=&V_{ub}^*\frac{G_F}{\sqrt{2}}\bigg[
\sum_{\l=0,s} \eta_\l\bar{L}^{\l_l}_\l (\Pi_{f_0} \overline{S}^0_\l Y^0_0 +
\xi^* \Pi_D \overline{P}^0_\l \tilde{Y}^0_0+\Pi_\rho \overline{V}^0_\l Y^0_1
+\Pi_{f_2} \overline{T}^0_\l Y^0_2)\nonumber\\
&& +\sum_{\l=\pm 1} \bar{L}^{\l_l}_\l (\Pi_\rho \overline{V}^\l_\l Y^\l_1
+\Pi_{f_2} \overline{T}^\l_\l Y^\l_2)\bigg],
\label{ampbar}
\end{eqnarray}
where
\begin{eqnarray}
&&\overline{S}^0_{\l_W}=S^0_{\l_W},\nonumber\\
&&\overline{P}^0_{\l_W}=-P^0_{\l_W},\nonumber\\
&&\overline{V}^0_{0,s}=V^0_{0,s},\; \overline{V}^{\pm 1}_{\pm 1}=V^{\mp 1}_{\mp 1},\nonumber\\
&&\overline{T}^0_{0,s}=T^0_{0,s},\; \overline{T}^{\pm 1}_{\pm 1}=T^{\mp 1}_{\mp 1}.
\end{eqnarray}
It is easy to see that
if $V_{ub}$ and $V_{cb}$ are real, the amplitude (\ref{amp}) of the $B^-$ decay and
(\ref{ampbar}) of the $B^+$ decay satisfy the CP relation:
\begin{eqnarray}
{\cal A}^{\pm}(\theta^*,\phi^*,\theta_l)
=\eta_{CP}{\bar{\cal A}}^{\mp}(\theta^*,-\phi^*,\theta_l;
\tilde{Y}^0_0\to-\tilde{Y}^0_0),
\label{cprel}
\end{eqnarray}
where $\theta^*$ and $\phi^*$ in $\bar{\cal A}^{\l_l}$ are the angles of
the final state $\pi^+$, while those in ${\cal A}^{\l_l}$ are for $\pi^-$.
Then, with a complex weak phase $\xi$,
$d\Gamma/d\Phi_4$ can be decomposed into a CP-even part ${\cal S}$ and
a CP-odd part ${\cal D}$:
\begin{eqnarray}
\frac{d\Gamma}{d\Phi_4}=\frac{1}{2}({\cal S}+{\cal D}).
\end{eqnarray}
The CP-even part ${\cal S}$ and the CP-odd part ${\cal D}$ can be easily
identified by making use of the CP relation (\ref{cprel}) between $B^-$ and
$B^+$ decay amplitudes, and they are expressed as
\begin{eqnarray}
{\cal S} =\frac{d(\Gamma+\overline{\Gamma})}{d\Phi_4},\quad
{\cal D} =\frac{d(\Gamma-\overline{\Gamma})}{d\Phi_4},
\end{eqnarray}
where $\Gamma$ and $\overline{\Gamma}$ are the decay rates for $B^-$ and $B^+$,
respectively, and
we have used the same kinematic variables $\{s_{_M},q^2,\theta^*,\theta_l\}$
for the $d\overline{\Gamma}/d\Phi_4$ except for the replacements of
${\phi^*} \to -\phi^*$ and $\tilde{Y}^0_0 \to -\tilde{Y}^0_0$,
as shown in Eq. (\ref{cprel}).
The CP-even ${\cal S}$ term and the CP-odd ${\cal D}$ term can be obtained from
$B^\mp$ decay probabilities, and their explicit form is listed in Appendix C.
Note that the CP-odd term is proportional to the imaginary part
of the parameter $\xi$ in Eq.~(\ref{xi}).
Before we go further on to the beyond the SM analyses,
we note that in addition to the resonant tree diagram contributions
there are other SM contributions through
annihilation diagrams and electroweak penguin diagrams,
which are relevant for nonresonant case.
As written in Section 1, we consider only resonant contributions by assuming
nonresonant contributions can be separated through data analyses.
\begin{center}
{\bf B. With complex scalar couplings}
\end{center}
Next we consider CP violation effects in extensions of the SM, where we extend
the virtual $W$-exchange part in Fig.~1 by including an additional scalar
interaction with complex couplings.
First we describe the formalism in a model independent way, but later
we consider specific models such as multi-Higgs doublet models and
scalar-leptoquark models.
In this case CP-violating phases can be generated through the interference
between $W$-exchange diagrams and scalar exchange diagrams with complex couplings.
The decay amplitudes for $B^- \to \pi^+ \pi^- l^- \bar{\nu}_l$ are expressed as
\begin{eqnarray}
{\cal A}^{\lambda_l}&=&-V_{ub}\frac{G_F}{\sqrt{2}}\frac{1}{\sqrt{2}}\sum_{i}\sum_{\lambda_i}
\bigg[\langle l^-(p_l,\lambda_l)\bar{\nu}(p_\nu)|j^{\mu\dagger}|0\rangle
\langle M_i(p_i,\lambda_i)|J_{\mu}|B^-(p_B)\rangle \nonumber\\
& &+\zeta\langle l^-(p_l,\lambda_l)\bar{\nu}(p_\nu)|j_s^{\dagger}|0\rangle
\langle M_i(p_i,\lambda_i)|J_s|B^-(p_B)\rangle \bigg]\nonumber\\
& &\times \Pi_i(s_{_M})\langle \pi^+(p_+)\pi^-(p_-)\|M_i(p_i,\lambda_i)\rangle ,
\end{eqnarray}
where
\begin{eqnarray}
&&j^\mu=\bar{\psi}_\nu\gamma^\mu(1-\gamma_5)\psi_l,\nonumber\\
&&J^\mu=\bar{\psi}_u\gamma^\mu(1-\gamma_5)\psi_b\equiv V^\mu-A^\mu,
\end{eqnarray}
and their corresponding scalar currents are
\begin{eqnarray}
j_s=\bar{\psi}_\nu(1-\gamma_5)\psi_l,\;\;
J_s=\bar{\psi}_u(1-\gamma_5)\psi_b,
\end{eqnarray}
the additional factor $1/\sqrt{2}$ comes from the isospin factor as mentioned earlier.
Here the parameter $\zeta$, which parameterizes contributions from physics beyond
the SM, is in general a complex number.
And as explained earlier,
in order to exclude any possible CP violation effects induced within the SM,
we only include the lowest three states $\rho(770)$, $f_0(980)$
and $f_2(1270)$ as intermediate states.
By using the Dirac equation for the leptonic current, $q_\mu j^\mu=m_l j_s$,
the amplitude can be written as
\begin{eqnarray}
{\cal A}^{\lambda_l}&=&-V_{ub}\frac{G_F}{2}\sum_{i}\sum_{\lambda_i}
\langle l^-(p_l,\lambda_l)\bar{\nu}(p_\nu)|j^{\mu\dagger}|0\rangle
\langle M_i(p_i,\lambda_i)|\Omega_\mu|B^-(p_B)\rangle \nonumber\\
& &\times \Pi_i(s_{_M})\langle \pi^+(p_+)\pi^-(p_-)\|M_i(p_i,\lambda_i)\rangle ,
\end{eqnarray}
where the effective hadronic current $\Omega_\mu$ is defined as
\begin{eqnarray}
\Omega_\mu\equiv J_\mu + \zeta \frac{q_\mu}{m_l} J_s.
\end{eqnarray}
In this case the amplitudes ${\cal M}^{\l_l}_{\l_i}$ of $B\to M_i l\bar{\nu}$
have the same form as the previous SM case (\ref{smamp}) except for the
modification in the hadronic current part due to the additional scalar current:
\begin{eqnarray}
{\cal M}^{\l_l}_{\l_i}=\frac{G_F}{2}V_{ub}
\sum_{\l_W}\eta_{\l_W}L^{\l_l}_{\l_W}{\cal H}^{\l_i}_{\l_W},
\end{eqnarray}
where ${\cal H}^{\l_i}_{\l_W}$ stands for the hadronic amplitudes modified
by the scalar current $J_s$.
Using the equation of motion for $u$ and $b$ quarks, we get within the on-shell
approximation
\begin{eqnarray}
J_s = (p_b^\mu-p_u^\mu) \left[\frac{V_\mu}{m_b-m_u}+\frac{A_\mu}{m_b+m_u}\right] .
\end{eqnarray}
Later for numerical calculations,
we use the approximation,
$(p_b^\mu-p_u^\mu) \approx (p_B^\mu-p_{M_i}^\mu) \equiv q^\mu$,
which is assumed in quark model calculations of form factors \cite{ISGW}.
After explicit calculation, we find that the additional scalar current modifies
only the scalar component of ${\cal H}^{\l_i}_{\l_W}$: {\it i.e.}
\begin{eqnarray}
{\cal H}^0_s&=&(1-\zeta^\prime) H^0_s,\nonumber\\
{\rm and}~~~{\cal H}^{\l_i}_{\l_W}&=&H^{\l_i}_{\l_W}
\;\;{\rm for}\;\; \l_W = 0,~\pm 1,
\end{eqnarray}
where
\begin{eqnarray}
\zeta^\prime=\frac{q^2}{m_l(m_b+m_u)}\zeta .
\label{zetap}
\end{eqnarray}
In this case,
$d\Gamma/d\Phi_4$ also can be decomposed into a CP-even part ${\cal S}$ and
a CP-odd part ${\cal D}$:
\begin{eqnarray}
\frac{d\Gamma}{d\Phi_4}=\frac{1}{2}({\cal S}+{\cal D}).
\end{eqnarray}
Their explicit form is listed in Appendix C.
Note that the CP-odd term is proportional to the imaginary part
of the parameter $\zeta$ and the lepton mass $m_l$.
Therefore, we have to consider massive leptonic ($\mu$ or $\tau$) decays.
As specific extensions of the SM,
we consider four types of scalar-exchange models which preserve the
symmetries of the SM \cite{scalar}:
One of them is the multi-Higgs-doublet (MHD) model \cite{Grossman} and
the other three models are scalar-leptoquark (SLQ) models \cite{wyler,randall}.
The authors of Ref.~\cite{tau} investigated CP violations in $\tau$ decay processes
with these extended models. We follow their description and
make it to be appropriate for our analysis.
In the MHD model CP violation can arise in the charged Higgs sector
with more than two Higgs doublets \cite{Wein} and when not all the charged
scalars are degenerate. As in most previous phenomenological analyses,
we assume that all but the lightest of the charged
scalars effectively decouple from fermions.
The effective Lagrangian of the MHD model contributing
to the decay $B\rightarrow \pi\pi l\bar{\nu}_l$
is then given at energies considerably low compared to $M_H$ by
\begin{eqnarray}
{\cal L}_{_{MHD}}=2\sqrt{2}G_FV_{ub}\frac{m_l}{M_H^2}\Big[m_b XZ^*(\bar{u}_L b_R)
+m_u YZ^*(\bar{u}_R b_L)\Big](\bar{l}_R \nu_L),
\end{eqnarray}
where $X$, $Y$ and $Z$ are complex coupling constants which can be
expressed in terms of the charged Higgs mixing matrix elements.
{}From the effective Lagrangian, we obtain for the MHD CP-violation parameter
${\rm Im}(\zeta_{_{MHD}})$,
\begin{eqnarray}
\Im(\zeta_{_{MHD}})=\frac{m_l m_b}{M_H^2}\Big\{\Im(XZ^*)-(\frac{m_u}{m_b})\Im(YZ^*)\Big\}.
\label{etaMHD}
\end{eqnarray}
The constraints on the CP-violation parameter (\ref{etaMHD}) depend upon
the values chosen for the $u$ and $b$ quark masses.
In the present work, we use (See Appendix A)
\begin{eqnarray}
m_u=0.33\;{\rm GeV},\qquad m_b=5.12\;{\rm GeV}.
\end{eqnarray}
In the MHD model the strongest constraint \cite{Grossman} on $\Im(XZ^*)$
comes from the measurement of the branching ratio
${\cal B}(b\to X\tau \nu_\tau)$, which actually gives a constraint on $|XZ|$.
For $M_H<440$ GeV, the bound on $\Im(XZ^*)$ is given by
\begin{eqnarray}
\Im(XZ^*)\;<\;|XZ|\;<0.23M^2_H \;{\rm GeV}^{-2}.
\end{eqnarray}
On the other hand, the bound on $\Im(YZ^*)$ is mainly
given by $K^+\to \pi^+\nu\bar{\nu}$. The present bound \cite{Grossman} is
\begin{eqnarray}
\Im(YZ^*)\;<\;|YZ|\;<110.
\end{eqnarray}
Combining the above bounds, we obtain
the following bounds on $\Im(\zeta_{_{MHD}})$ as
\begin{eqnarray}
&&|\Im(\zeta_{_{MHD}})|\;<\; 2.06\;\qquad\qquad {\rm for}\;\tau\;{\rm family},\nonumber\\
&&|\Im(\zeta_{_{MHD}})|\;<\; 0.12\;\qquad\qquad {\rm for}\;\mu\;{\rm family}.
\label{Hmodel}
\end{eqnarray}
On the other hand, the effective Lagrangians for the three SLQ models \cite{scalar,wyler}
contributing to the decay $B\to \pi\pi l\nu$ are written in the form,
after a few Fierz rearrangements:
\begin{eqnarray}
{\cal L}^{^{I}}_{_{SLQ}}&=&-\frac{x_{3j}x^{\prime *}_{1j}}{2M^2_{\phi_1}}\left[
(\bar{b}_Lu_R)(\bar{\nu}_{lL}l_R)
+\frac{1}{4}(\bar{b}_L\sigma^{\mu\nu}u_R)
(\bar{\nu}_{lL}\sigma_{\mu\nu}l_R)\right]+h.c.,\nonumber\\
{\cal L}^{^{II}}_{_{SLQ}}&=&-\frac{y_{3j}y^{\prime *}_{1j}}{2M^2_{\phi_2}}\left[
(\bar{b}_Lu_R)(\bar{l}^c_R\nu^c_{l L})
+\frac{1}{4}(\bar{b}_L\sigma^{\mu\nu}u_R)
(\bar{l}^c_R\sigma_{\mu\nu}\nu^c_{l L})\right]\nonumber\\
&&+\frac{y_{3j}y^*_{1j}}{2M^2_{\phi_2}}(\bar{b}_L\gamma_\mu u_L)
(\bar{l}^c_L\gamma^\mu\nu^c_{l L})+h.c.,\nonumber\\
{\cal L}^{^{III}}_{_{SLQ}}&=&-\frac{z_{3j}z^*_{1j}}{2M^2_{\phi_3}}(\bar{b}_L\gamma_\mu u_L)
(\bar{l}^c_L\gamma^\mu\nu^c_{l L})+h.c.\;,
\end{eqnarray}
where $j=2,3$ for $l=\mu,\tau$, respectively
and the coupling constants $x^{(\prime)}_{ij}$, $y^{(\prime)}_{ij}$ and
$z_{ij}$ are in general complex so that CP is violated
in the scalar-fermion Yukawa interaction terms. The superscript $c$ in the
Lagrangians ${\cal L}^{^{II}}_{_{SLQ}}$ and ${\cal L}^{^{III}}_{_{SLQ}}$ denotes
charge conjugation, {\it i.e.} $\psi^c_{R,L}=i\gamma^0\gamma^2\bar{\psi}^T_{R,L}$
in the chiral representation.
Then we find that the size of the SLQ model
CP-violation effects is dictated by the CP-odd parameters
\begin{eqnarray}
&&{\rm Im}(\zeta^{^I}_{_{SLQ}})=-\frac{{\rm Im}[x_{3j}
x^{\prime *}_{1j}]}{4\sqrt{2}G_FV_{ub}M^2_{\phi_1}}\,,\nonumber\\
&&{\rm Im}(\zeta^{^{II}}_{_{SLQ}})= -\frac{{\rm Im}[y_{3j}y^{\prime *}_{1j}]}
{4\sqrt{2}G_FV_{ub}M^2_{\phi_2}}\,,\nonumber\\
&&{\rm Im}(\zeta^{^{III}}_{_{SLQ}})= 0\,.
\label{LQp}
\end{eqnarray}
Although there are at present no direct
constraints on the SLQ model CP-odd parameters in (\ref{LQp}),
a rough constraint to the parameters can be provided by the assumption
\cite{Davidson} that $|x^\prime_{1j}|\sim |x_{1j}|$ and
$|y^\prime_{1j}|\sim |y_{1j}|$, that is to say, the leptoquark couplings
to quarks and leptons belonging to the same generation are of a similar
size; then the experimental upper bounds
from $B\bar{B}$ mixing for $\tau$ family, $B\to \mu\bar{\mu}X$ decay
for $\mu$ in Model I, and $B\to l\nu X$ for $\tau$
together with the $V_{ub}$ measurement
for $\mu$ in Model II yield \cite{Davidson}
\begin{eqnarray}
&&|{\rm Im}(\zeta^{^I}_{_{SLQ}})|< 2.76,\quad
|{\rm Im}(\zeta^{^{II}}_{_{SLQ}})|< 18.4 \qquad\; {\rm for}\;\tau\;{\rm family},\nonumber\\
&&|{\rm Im}(\zeta^{^I}_{_{SLQ}})|< 0.37,\quad
|{\rm Im}(\zeta^{^{II}}_{_{SLQ}})|< 1.84 \qquad\; {\rm for}\;\mu\;{\rm family}.
\label{Models}
\end{eqnarray}
Based on the constraints (\ref{Hmodel}) and (\ref{Models}) to the CP-odd
parameters, we quantitatively estimate the number of
$B^-\to \pi^+\pi^- l^-\bar{\nu}_l$ decays to detect CP violation
for the maximally-allowed values of the CP-odd parameters.
\section{Observable CP Asymmetries}
\noindent
An easily constructed CP-odd asymmetry is the rate asymmetry
\begin{eqnarray}
A\equiv \frac{\Gamma-\overline{\Gamma}}{\Gamma+\overline{\Gamma}},
\end{eqnarray}
which has been used as a
probe of CP violation in Higgs and top quark sectors \cite{rateasym}.
Here $\Gamma$ and $\overline{\Gamma}$ are the decay rates for $B^-$ and $B^+$,
respectively.
The statistical significance of the asymmetry can be computed as
\begin{eqnarray}
N_{SD}=\frac{N_- -N_+}{\sqrt{N_-+N_+}} =\frac{N_- -N_+}{\sqrt{N \cdot Br}},
\end{eqnarray}
where $N_{SD}$ is the number of standard deviations,
$N_\pm$ is the number of events predicted in $B_{l4}$ decay for $B^\pm$
meson, $N$ is the number of $B$-mesons produced,
and $Br$ is the branching fraction of the relevant $B$ decay mode.
For a realistic detection efficiency
$\epsilon$, we have to rescale the number of events by this parameter,
$N_-+N_+\to\epsilon (N_-+N_+)$. Taking $N_{SD}=1$, we obtain
the number $N_B$ of the B mesons needed to observe CP violation at $1$-$\sigma$ level:
\begin{eqnarray}
N_B=\frac{1}{Br\cdot A^2}.
\end{eqnarray}
Next, we consider the so-called optimal observable.
An appropriate real weight function $w(s_{_M},q^2;\theta^*,\theta_l,\phi^*)$
is usually employed to separate the CP-odd ${\cal D}$ contribution and to enhance
its analysis power for the CP-odd parameter through
the CP-odd quantity:
\begin{eqnarray}
\langle w{\cal D}\rangle\equiv\int\left[w{\cal D}\right] d\Phi_4,
\end{eqnarray}
and the analysis power is determined by the parameter,
\begin{eqnarray}
\varepsilon
=\frac{\langle w{\cal D}\rangle}{\sqrt{\langle{\cal S}\rangle
\langle w^2{\cal S}\rangle}}\;.
\label{Significance}
\end{eqnarray}
For the analysis power $\varepsilon$, the number $N_B$ of the $B$-mesons
needed to observe CP violation at 1-$\sigma$ level is
\begin{eqnarray}
N_B=\frac{1}{Br\cdot\varepsilon^2}\;.
\label{eq:number}
\end{eqnarray}
Certainly, it is desirable to find the optimal weight function
with the largest analysis power. It is known \cite{Optimal} that
when the CP-odd contribution to the total rate is relatively small,
the optimal weight function is approximately given as
\begin{eqnarray}
w_{\rm opt}(s_{_M},s_L;\theta,\theta_l,\phi)=
\frac{{\cal D}}{{\cal S}}~~~\Rightarrow~~~
\varepsilon_{\rm opt}=\sqrt{\frac{\langle\frac{{\cal D}^2}{{\cal S}}\rangle}
{\langle{\cal S}\rangle}}.
\end{eqnarray}
We adopt this optimal weight function in the following numerical analyses.
\section{Numerical Results and Conclusions}
\noindent
Now we show our numerical results.
We use the so-called ISGW predictions \cite{ISGW}
for all the form factors in $B\to M_i$ transition amplitudes of Eq.~(\ref{BMamp}).
One can find in Ref.~\cite{ISGW} the detailed description of the general
formalism and relevant form factors for $B\to Xe\bar{\nu}_e$
after neglecting lepton masses.
In Appendix A, we give explicit expressions of form factors needed for
semileptonic decays with non-zero lepton masses.
\begin{table}
{Table~2}. {The CP-violating rate asymmetry A and the optimal asymmetry
$\varepsilon_{\rm opt}$, determined within the SM,
and the number of charged B meson pairs,
$N_B$, needed for detection at $1\sigma$ level,
at reference value $\Im(\xi)=12.5$.}\par
\begin{tabular}{c|cc|cc|cc}
\hline
\multicolumn{7}{c}{$B\to \pi^+\pi^-l\bar{\nu}_l$}\\
\hline
Mode
&\multicolumn{2}{c|}{$l=e$}&\multicolumn{2}{c|}{$l=\mu$}&\multicolumn{2}{c}{$l=\tau$}\\
\hline\hline
Asym. & Size(\%) & $N_B$ & Size(\%) & $N_B$ & Size(\%) & $N_B$\\
\hline
A & $0.94\times 10^{-6}$ & $1.37\times 10^{18}$
& $1.71\times 10^{-6}$ & $4.16\times 10^{17}$ & $1.14\times 10^{-6}$ & $1.46\times 10^{18}$\\
$\varepsilon_{\rm opt}$ & $1.45\times 10^{-2}$ & $5.75\times 10^{9}$
& $1.44\times 10^{-2}$ & $5.79\times 10^{9}$ & $1.11\times 10^{-2}$ & $1.56\times 10^{10}$\\
\hline
\end{tabular}
\end{table}
We first consider the case within the SM.
Total branching ratio of the $B^-\to (\sum_i M_i \to \pi^+\pi^-) e\bar{\nu}_e,~
M_i=\rho, f_{0,2}, D$ is about $0.8\%$.
It depends on the chosen value for $|V_{ub}|$, and
here we adopt the result by CLEO \cite{CLEO}
\begin{eqnarray}
|V_{ub}|=3.3\pm 0.4 \pm 0.7\times 10^{-3}.
\end{eqnarray}
In Table 2, we show the results of $B\to \pi\pi l\nu$ decays
for the two CP-violating asymmetries;
the rate asymmetry $A$ and the optimal asymmetry $\varepsilon_{\rm opt}$.
We estimated the number of $B$-meson pairs, $N_B$, needed for detection at
$1\sigma$ level for maximally-allowed values of CP-odd parameters $\Im(\xi)$ in
Eq.~(\ref{xi}). We use the current experimental bound \cite{pdg}
\begin{eqnarray}
\left| \frac{V_{ub}}{V_{cb}}\right|=0.08\pm 0.02,
\end{eqnarray}
which means
\begin{eqnarray}
|\xi|=12.5\pm 3.13.
\end{eqnarray}
The results in Table 2 are for the maximal case with $\Im(\xi)=|\xi|=12.5$.
Due to the large cancelations in the simple rate asymmetry \cite{cancel}
when we integrated over the phase space,
the optimal observable gives much better result.
For example, using the optimal observable, we need $\sim 10^9$
$B$-meson pairs to detect the maximal CP-odd effect in electron mode.
CP violation effects in $B\to \pi^+\pi^-l\bar{\nu}_l$ decays within the SM
are not likely to be detected, with ${\cal O}(10^8)$ $B$-meson pairs to be produced
at the asymmetric $B$ factories.
One may rely on hadronic $B$-factories of BTeV and LHC-B.
\begin{table}
{Table~3}. {The CP-violating rate asymmetry A and the optimal asymmetry
$\varepsilon_{\rm opt}$, determined in the extended models,
and the number of charged B meson pairs,
$N_B$, needed for detection at $1\sigma$ level,
at reference values
(a) $\Im(\zeta_{_{MHD}})=2.06$, $\Im(\zeta^{^I}_{_{SLQ}})=2.76$ and
$\Im(\zeta^{^{II}}_{_{SLQ}})=18.4$ for the $B_{\tau 4}$ decays
and (b) $\Im(\zeta_{_{MHD}})=0.12$, $\Im(\zeta^{^I}_{_{SLQ}})=0.37$ and
$\Im(\zeta^{^{II}}_{_{SLQ}})=1.84$ for the $B_{\mu 4}$ decays.}\par
\begin{tabular}{c|cc|cc|cc}
\hline
\multicolumn{7}{c}{(a) $B^-\to \pi^+\pi^-\tau\bar{\nu}_\tau$ mode}\\
\hline
Model
&\multicolumn{2}{c|}{MHD}&\multicolumn{2}{c|}{SLQ I}&\multicolumn{2}{c}{SLQ II}\\
\hline\hline
Asym. & Size(\%) & $N_B$ & Size(\%) & $N_B$ & Size(\%) & $N_B$\\
\hline
A & $1.47\times 10^{-3}$ & $7.63\times 10^{11}$ & $2.67\times 10^{-3}$ & $1.99\times 10^{11}$
& $3.62\times 10^{-3}$ & $7.39\times 10^{9}$\\
$\varepsilon_{\rm opt}$ & $16.2$ & $6.23\times 10^3$ & $18.2$ &
$4.27\times 10^3$ & $9.67$ & $1.04\times 10^{3}$\\
\hline
\multicolumn{7}{c}{ }\\
\hline
\multicolumn{7}{c}{(b) $B^-\to \pi^+\pi^-\mu\bar{\nu}_\mu$ mode}\\
\hline
Model
&\multicolumn{2}{c|}{MHD}&\multicolumn{2}{c|}{SLQ I}&\multicolumn{2}{c}{SLQ II}\\
\hline\hline
Asym. & Size(\%) & $N_B$ & Size(\%) & $N_B$ & Size(\%) & $N_B$\\
\hline
A & $2.61\times 10^{-5}$ & $1.93\times 10^{15}$ & $0.90\times 10^{-4}$ & $1.59\times 10^{14}$
& $3.43\times 10^{-4}$ & $8.89\times 10^{12}$\\
$\varepsilon_{\rm opt}$ & $0.18$ & $3.89\times 10^7$ & $0.50$ &
$5.13\times 10^6$ & $1.48$ & $4.76\times 10^{5}$\\
\hline
\end{tabular}
\end{table}
Next we consider the extended model case.
In this case, CP violation effects are proportional to the lepton mass,
and we consider only massive lepton ($\mu$ or $\tau$) cases.
In Table 3, we show the results of $B_{\tau 4}$ and $B_{\mu 4}$ decays.
Here in order to distinguish new physics effect from the SM one,
we use a cutoff for the invariant mass of the final state $\pi^+\pi^-$ as
\begin{eqnarray}
\sqrt{s_{_M}}\le 1.4\;{\rm GeV}.
\end{eqnarray}
We consider only the lowest three
$u\bar{u}$ states, $\rho(770)$, $f_0(980)$ and $f_2(1275)$
as intermediate resonances in Table 1
so that the effects of $D$ meson can not enter, and we can thus ensure
that the result is solely from new physics.
Similarly as in the SM case,
we estimate the number of $B$-meson pairs, $N_B$, needed for detection at
$1\sigma$ level for {\it maximally-allowed} values of CP-odd parameters $\Im(\zeta)$ of
Eq.~(\ref{Hmodel}) and (\ref{Models}).
We again find the optimal observable gives much better results than
the simple rate asymmetry.
The results in Table 3 show that CP violation effects from new physics
are readily observed in the forthcoming asymmetric $B$-factories,
by using optimal observables.
As expected,
$B_{\tau 4}$ decay modes give better results than $B_{\mu 4}$ cases
for the MHD model, where the CP-odd parameter itself is proportional to
the lepton mass.
For example, the current bounds in the MHD model
$$\Im(\zeta_{_{MHD}})=2.06\;(0.12)~~~{\rm for}~~~ \tau\;(\mu)$$
directly result from the lepton mass dependence.
But there is no such dependence in the SLQ models.
The current numerical values of CP-odd parameters in the SLQ models,
\begin{eqnarray}
\Im(\zeta^{^I}_{_{SLQ}})&=&2.76\;(0.37) \nonumber\\
\Im(\zeta^{^{II}}_{_{SLQ}})&=&18.4\;(1.84)~~~{\rm for}~~~ \tau\;(\mu), \nonumber
\end{eqnarray}
are just from different experimental bounds.
Therefore, the smaller CP-odd value for $\mu$ family is a consequence of the fact
that the current experimental constraints on the muon mode are more stringent.
And $B_{\tau 4}$ decay modes would provide more stringent constraints
to all the extended models that we have considered.
In conclusion, we have investigated
direct CP violations from physics beyond the SM as well as within the SM
through semileptonic $B_{l4}$
decays: $B^\pm\to \pi^+\pi^- l^\pm\nu_l$.
Within the SM, CP violation could be generated through interference between
resonances with different quark flavors, that is, with different CKM matrix
elements.
We included $u\bar{u}$ state mesons $(\rho, f_0\;{\rm and}\;f_2)$ and
$D$ meson as intermediate resonances which decay to $\pi^+\pi^-$.
Using optimal observables, we found ${\cal O}(10^9)$ $B$-meson pairs are
needed to probe CP violation effects at $1\sigma$ level
for the current maximal value of $\Im(\xi)=|V_{cb}/V_{ub}|=12.5$.
We have also investigated CP violation effects in extensions of the SM.
We considered multi-Higgs doublet model and scalar-leptoquark models.
Here CP violation is implemented
through interference between $W$-exchange diagrams and
scalar-exchange diagrams with complex couplings in the extended models.
We calculated the CP-odd rate asymmetry and the optimal asymmetry
for $B_{\tau 4}$ and $B_{\mu 4}$ decay modes.
We found that the optimal asymmetries for both modes are sizable and
can be detected at $1\sigma$ level with about $10^3$-$10^7$
$B$-meson pairs,
for maximally-allowed values of CP-odd parameters.
Since $\sim$$10^8$ $B$-meson pairs are expected to be produced yearly
at the asymmetric $B$ factories,
one could easily investigate CP-violation effects in these decay modes $B_{l4}$
to extract much more stringent constraints on CP-odd parameters,
$\Im(\zeta_{_{MHD}})$ and $\Im(\zeta^{^{I,II}}_{_{SLQ}})$.
\section*{Acknowledgments}
\noindent
We thank G. Cvetic for careful reading of the manuscript and
his valuable comments.
The work of C.S.K. was supported
in part by KRF Non-Directed-Research-Fund, Project No. 1997-001-D00111,
in part by the BSRI Program, Ministry of Education, Project No. 98-015-D00061,
in part by the KOSEF-DFG large collaboration project,
Project No. 96-0702-01-01-2.
J.L. and W.N. wish to acknowledge the financial support of
1997-sughak program of Korean Research Foundation.
\newpage
\noindent
{\Large\bf Appendix}
\begin{appendix}
\section{Form factors}
Form factors in Eq.~(\ref{BMamp}) within ISGW model \cite{ISGW} are
\begin{eqnarray}
u_+(q^2)=-F_5(q^2;f_0)\frac{m_u m_b m_q}{\sqrt{6}\beta_B \tilde{m}_{f_0}\mu_-},\;\;
u_+(q^2)=F_5(q^2;f_0)\frac{m_u(\tilde{m}_B+\tilde{m}_{f_0})}{\sqrt{6}\beta_B\tilde{m}_{f_0}}
\end{eqnarray}
\begin{eqnarray}
&&g(q^2)=\frac{F_3(q^2;\rho)}{2}\left[\frac{1}{m_q}-\frac{m_u \beta^2_B}
{2\mu_-\tilde{m}_\rho \beta^2_{B\rho}}\right],\;\;
f(q^2)=2\tilde{m}_B F_3(q^2;\rho),\nonumber\\
&&a_+(q^2)=-\frac{F_3(q^2;\rho)}{2\tilde{m}_\rho}\left[1+\frac{m_u}{m_b}
\left(\frac{\beta_B^2-\beta_\rho^2}{\beta_B^2+\beta_\rho^2}\right)
-\frac{m_u^2\beta_\rho^2}{4\mu_-\tilde{m}_B\beta_{B\rho}^4}\right],\nonumber\\
&&a_-(q^2)=\frac{F_3(q^2;\rho)}{2\tilde{m}_\rho}\left[1+\frac{m_u}{m_b}
\left(1+\frac{m_u\beta_\rho^2}{m_q\beta_{B\rho^2}}\right)
-\frac{m_u^2\beta_\rho^2}{4\mu_+\tilde{m}_B\beta_{B\rho}^4}\right]
\end{eqnarray}
\begin{eqnarray}
&&h(q^2)=F_5(q^2;f_2)\frac{m_u}{2\sqrt{2}\tilde{m}_B\beta_B}
\left[\frac{1}{m_q}-\frac{m_u \beta^2_B}
{2\mu_-\tilde{m}_{f_2} \beta^2_{Bf_2}}\right],\;\;
k(q^2)=\sqrt{2}\frac{m_u}{\beta_B} F_5(q^2;f_2),\nonumber\\
&&b_+(q^2)=-\frac{F_5(q^2;f_2)m_u}{2\sqrt{2}\beta_B\tilde{m}_{f_2}m_b}\left[1
-\frac{m_u\beta_{f_2}^2}{\tilde{m}_B\beta_{Bf_2}^2}
-\frac{m_u^2m_b\beta_{f_2}^4}{4\mu_-\tilde{m}_B^2\beta_{Bf_2}^4}\right],\nonumber\\
&&b_-(q^2)=\frac{F_5(q^2;f_2)m_u}{2\sqrt{2}\beta_B\tilde{m}_{f_2}m_b}\left[1
+\frac{m_u^2\beta_{f_2}^2}{m_q\tilde{m}_B\beta_{Bf_2}^2}
-\frac{m_u^2m_b\beta_{f_2}^4}{4\mu_+\tilde{m}_B^2\beta_{Bf_2}^4}\right]
\end{eqnarray}
\begin{eqnarray}
&&f_+(q^2)=F_3(q^2;D)\left[1+\frac{m_b}{2\mu_-}-\frac{m_bm_qm_u\beta_B^2}
{4\mu_+\mu_-\tilde{m}_D\beta_{BD}^2}\right],\nonumber\\
&&f_-(q^2)=F_3(q^2;D)\left[1-(\tilde{m}_B+\tilde{m}_D)
\left(\frac{1}{2m_q}-\frac{m_u\beta_B^2}{4\mu_+\tilde{m}_D\beta_{BD}^2}\right)
\right]
\end{eqnarray}
where
\begin{eqnarray}
\beta_{BX}^2=\frac{1}{2}(\beta_B^2+\beta_X^2),\;\;
\mu_\pm=\left(\frac{1}{m_q}\pm\frac{1}{m_b}\right)^{-1}.
\end{eqnarray}
And
\begin{eqnarray}
F_n(q^2;X)=\left(\frac{\tilde{m}_X}{\tilde{m}_B}\right)^{1/2}
\left(\frac{\beta_B\beta_X}{\beta_{BX}^2}\right)^{n/2}
\exp\left[-\left(\frac{m_u^2}{4\tilde{m}_B\tilde{m}_X}\right)
\frac{q_m-q^2}{\kappa^2\beta_{BX}^2}\right],
\end{eqnarray}
where relativistic compensation factor $\kappa=0.7$, and $q_m$ is the maximum value
of $q^2$:
\begin{eqnarray}
q_m=(m_B-\sqrt{s_{_M}})^2,
\end{eqnarray}
and $m_q$ is $m_u$ for $u\bar{u}$ state mesons and $m_c$ for $D$-mesons.
The numerical values of $\beta_X$ in GeV unit are
\begin{eqnarray}
\beta_B=0.41,\;\beta_D=0.39,\;\beta_{f_0}=0.27,\;\beta_\rho=0.31,\;\beta_{f_2}=0.27,
\end{eqnarray}
and quark masses in GeV unit are
\begin{eqnarray}
m_u=0.33,\; m_c=1.82,\;m_b=5.12.
\end{eqnarray}
The so-called mock meson masses $\tilde{m}_X$ are defined as
\begin{eqnarray}
\tilde{m}_B=m_b+m_u,\; \tilde{m}_D=m_c+m_u,\; \tilde{m}_{\rho,f_0,f_2}=2m_u.
\end{eqnarray}
\section{Kinematics}
\begin{center}
{\bf Spherical harmonics}
\end{center}
\begin{eqnarray}
&&Y^0_0=\tilde{Y}^0_0=\frac{1}{\sqrt{4\pi}},\nonumber\\
&&Y_1^0=\sqrt{\frac{3}{4\pi}}\cos\theta,\;
Y_1^{\pm 1}=\mp\sqrt{\frac{3}{8\pi}}\sin\theta e^{\pm i\phi},\nonumber\\
&&Y_2^0=\sqrt{\frac{5}{4\pi}}(\frac{3}{2}\cos^2\theta-\frac{1}{2}),\;
Y_2^{\pm 1}=\mp\sqrt{\frac{15}{8\pi}}\sin\theta\cos\theta e^{\pm i\phi},
\end{eqnarray}
\begin{center}
{\bf Polarization vectors}
\end{center}
\noindent
In the $B$ rest frame, where the coordinates are chosen such that the $z$-axis is along the
$M_i$ momentum and the charged lepton momentum is in the $x$--$z$ plane with positive
$x$-component (cf. Fig.~2a),
the polarization vectors for the virtual $W$ are
\begin{eqnarray}
\epsilon(q,\pm)^\mu&=&\mp\frac{1}{\sqrt{2}}(0,1,\mp i,0),\nonumber\\
\epsilon(q,0)^\mu&=&\frac{1}{\sqrt{q^2}}(p_M,0,0,-q^0),\nonumber\\
\epsilon(q,s)^\mu&=&\frac{1}{\sqrt{q^2}}q^\mu,
\end{eqnarray}
and the polarization states of the spin 1 mesons are
\begin{eqnarray}
\epsilon(\pm 1)^\mu=\mp\frac{1}{\sqrt{2}}(0,1,\pm i,0),\;
\epsilon(0)^\mu=\frac{1}{\sqrt{s_{_M}}}(p_M,0,0,E_M),
\end{eqnarray}
where $p_M=\sqrt{Q_+Q_-}/2m_B$ with $Q_\pm$ defined in Eq.~(\ref{Qpm}),
and $E_M=(m_B^2+s_{_M}-q^2)/2m_B$.
And for the spin 2 meson we get
\begin{eqnarray}
\epsilon(\pm 2)^{\mu\nu}&=&\epsilon^\mu(\pm 1)\epsilon^\nu(\pm 1),\nonumber\\
\epsilon(\pm 1)^{\mu\nu}&=&\frac{1}{\sqrt{2}}\left[\epsilon^\mu(\pm 1)\epsilon^\nu(0)
+\epsilon^\mu(0)\epsilon^\nu(\pm 1)\right],\nonumber\\
\epsilon(0)^{\mu\nu}&=&\frac{1}{\sqrt{6}}\left[\epsilon^\mu(+1)\epsilon^\nu(-1)
+\epsilon^\mu(-1)\epsilon^\nu(+1)\right]+\sqrt{\frac{2}{3}}\epsilon^\mu(0)\epsilon^\nu(0).
\end{eqnarray}
In the $W$ rest frame the polarization states of the virtual $W$ are
\begin{eqnarray}
\epsilon(q,\pm)^\mu&=&\mp\frac{1}{\sqrt{2}}(0,1,\mp i,0),\nonumber\\
\epsilon(q,0)^\mu&=&(0,0,0,-1),\nonumber\\
\epsilon(q,s)^\mu&=&\frac{1}{\sqrt{q^2}}q^\mu=(1,0,0,0).
\end{eqnarray}
\section{CP-even and CP-odd quantities}
\begin{center}
{\bf Within the SM}
\end{center}
\noindent
The CP-even quantity ${\cal S}$ is
\begin{eqnarray}
{\cal S}=2C(q^2,s_{_M})\Sigma,
\end{eqnarray}
with
\begin{eqnarray}
\Sigma&=&(L^-_0S^0_0Y^0_0)^2|\Pi_{f_0}|^2+(L^-_0P^0_0Y^0_0)^2|\xi|^2|\Pi_D|^2+|\langle V^-\rangle \Pi_\rho|^2+|\langle T^-\rangle \Pi_{f_2}|^2\nonumber\\
&&+2(L^-_0S^0_0Y^0_0)\Re(\Pi_{f_0}\Pi_\rho^*\langle V^-\rangle^*+\Pi_{f_0}\Pi_{f_2}\langle T^-\rangle^*)+2\Re(\Pi_\rho\Pi_{f_2}^*\langle V^-\rangle \langle T^-\rangle^*)\nonumber\\
&&+2(L^-_0P^0_0Y^0_0)\Re(\xi)[(L^-_0S^0_0Y^0_0)\Re(\Pi_D\Pi_{f_0}^*)+\Re(\Pi_D\Pi_\rho^*\langle V^-\rangle^*+\Pi_D\Pi_{f_2}^*\langle T^-\rangle^*)]\nonumber\\
&&+(L^+_0S^0_0Y^0_0-L^+_sS^0_sY^0_0)^2|\Pi_{f_0}|^2+(L^+_0P^0_0Y^0_0-L^+_sP^0_sY^0_0)^2|\xi|^2|\Pi_D|^2\nonumber\\
&&+|\Pi_\rho|^2|\langle V^+\rangle -L^+_sV^0_sY^0_1|^2+|\Pi_{f_2}|^2|\langle T^+\rangle -L^+_sT^0_sY^0_2|^2\nonumber\\
&&+2(L^+_0S^0_0-L^+_sS^0_s)Y^0_0[-(L^+_sV^0_sY^0_1)\Re(\Pi_{f_0}\Pi_\rho^*)+\Re(\Pi_{f_0}\Pi_\rho^*\langle V^+\rangle^*)\nonumber\\
&& -(L^+_sT^0_sY^0_2)\Re(\Pi_{f_0}\Pi_{f_2}^*)+\Re(\Pi_{f_0}\Pi_{f_2}^*\langle T^+\rangle^*)]\nonumber\\
&&+2\Re(\Pi_\rho\Pi_{f_2}^*\langle V^+\rangle \langle T^+\rangle^*)+2(L^+_sV^0_sY^0_1)(L^+_sT^0_sY^0_2)\Re(\Pi_\rho\Pi_{f_2}^*)\nonumber\\
&&-2(L^+_sT^0_sY^0_2)\Re(\Pi_\rho\Pi_{f_2}^*\langle V^+\rangle )-2(L^+_sV^0_sY^0_1)\Re(\Pi_\rho\Pi_{f_2}^*\langle T^+\rangle^*)\nonumber\\
&&+2(L^+_0P^0_0-L^+_sP^0_s)Y^0_0\Re(\xi)[(L^+_0S^0_0-L^+_sS^0_s)Y^0_0\Re(\Pi_D\Pi_{f_0}^*)
-(L^+_sV^0_sY^0_1)\Re(\Pi_D\Pi_\rho^*)\nonumber\\
&& +\Re(\Pi_D\Pi_\rho^*\langle V^+\rangle^*)-(L^+_sT^0_sY^0_2)\Re(\Pi_D\Pi_{f_2}^*)+\Re(\Pi_D\Pi_{f_2}^*\langle T^+\rangle^*)],
\end{eqnarray}
and the CP-odd quantity ${\cal D}$ is
\begin{eqnarray}
{\cal D}=2\Im(\xi)C(q^2,s_{_M})\Delta,
\end{eqnarray}
with
\begin{eqnarray}
\Delta&=&-2(L^-_0P^0_0Y^0_0)[(L^-_0S^0_0Y^0_0)\Im(\Pi_D\Pi_{f_0}^*)+\Im(\Pi_D\Pi_\rho^*\langle V^-\rangle^*)+\Im(\Pi_D\Pi_{f_2}^*\langle T^-\rangle^*)]\nonumber\\
&&-2(L^+_0P^0_0-L^+_sP^0_s)Y^0_0[(L^+_0S^0_0-L^+_sS^0_s)Y^0_0\Im(\Pi_D\Pi_{f_0}^*)-(L^+_sV^0_sY^0_1)\Im(\Pi_D\Pi_\rho^*)\nonumber\\
&&+\Im(\Pi_D\Pi_\rho^*\langle V^+\rangle^*)-(L^+_sT^0_sY^0_2)\Im(\Pi_D\Pi_{f_2}^*)+\Im(\Pi_D\Pi_{f_2}^*\langle T^+\rangle^*)],
\end{eqnarray}
where
\begin{eqnarray}
\langle V^\pm\rangle \equiv\sum_{i=0,\pm 1}L^\pm_\l V^\l_\l Y^\l_1,\;\;
\langle T^\pm\rangle \equiv\sum_{i=0,\pm 1}L^\pm_\l T^\l_\l Y^\l_2,
\end{eqnarray}
and the overall function $C(q^2,s_{_M})$ is given by
\begin{eqnarray}
C(q^2,s_{_M})=|V_{ub}|^2\frac{G_F^2}{2}\frac{1}{2m_B}
\frac{(q^2-m_l^2)\sqrt{Q_+Q_-}}{256\pi^3m_B^2q^2}.
\end{eqnarray}
\begin{center}
{\bf With a complex scalar coupling}
\end{center}
\noindent
The CP-even quantity ${\cal S}$ is
\begin{eqnarray}
{\cal S}=2C(q^2,s_{_M})\Sigma,
\end{eqnarray}
with
\begin{eqnarray}
\Sigma&=&(L^-_0S^0_0Y^0_0)^2|\Pi_{f_0}|^2+|\langle V^-\rangle \Pi_\rho|^2+|\langle T^-\rangle \Pi_{f_2}|^2\nonumber\\
&&+2(L^-_0S^0_0Y^0_0)\Re(\Pi_{f_0}\Pi_\rho^*\langle V^-\rangle^*+\Pi_{f_0}\Pi_{f_2}^*\langle T^-\rangle^*)+2\Re(\Pi_\rho\Pi_{f_2}^*\langle V^-\rangle \langle T^-\rangle^*)\nonumber\\
&&+|\Pi_{f_0}|^2|L^+_0S^0_0Y^0_0-(1-\zeta^\prime)L^+_sS^0_sY^0_0|^2\nonumber\\
&&+|\Pi_\rho|^2[|\langle V^+\rangle|^2 +(L^+_sV^0_sY^0_1)^2|1-\zeta^\prime|^2
-2(L^+_sV^0_sY^0_1)\Re(\langle V^+\rangle)\Re(1-\zeta^\prime)]\nonumber\\
&& +|\Pi_{f_2}|^2[|\langle T^+\rangle|^2 +(L^+_sT^0_sY^0_2)^2|1-\zeta^\prime|^2
-2(L^+_sT^0_sY^0_2)\Re(\langle T^+\rangle)\Re(1-\zeta^\prime)]\nonumber\\
&&+2\Re(\Pi_{f_0}\Pi_\rho^*)[(L^+_0S^0_0-L^+_sS^0_s)Y^0_0\Re(\langle V^+\rangle )-(L^+_0S^0_0Y^0_0)(L^+_sV^0_sY^0_1)\Re(1-\zeta^\prime)\nonumber\\
&& +(L^+_sS^0_sY^0_0)\Re(\langle V^+\rangle )\Re(\zeta^\prime)+(L^+_sS^0_sY^0_0)(L^+_sV^0_sY^0_1)|1-\zeta^\prime|^2]\nonumber\\
&&+2\Im(\Pi_{f_0}\Pi_\rho^*)\Im(\langle V^+\rangle )[(L^+_0S^0_0-L^+_sS^0_s)Y^0_0+(L^+_sS^0_sY^0_0)\Re(\zeta^\prime)]\nonumber\\
&&+2\Re(\Pi_{f_0}\Pi_{f_2}^*)[(L^+_0S^0_0-L^+_sS^0_s)Y^0_0\Re(\langle T^+\rangle )-(L^+_0S^0_0Y^0_0)(L^+_sT^0_sY^0_2)\Re(1-\zeta^\prime)\nonumber\\
&& +(L^+_sS^0_sY^0_0)\Re(\langle T^+\rangle )\Re(\zeta^\prime)+(L^+_sS^0_sY^0_0)(L^+_sT^0_sY^0_2)|1-\zeta^\prime|^2]\nonumber\\
&&+2\Im(\Pi_{f_0}\Pi_{f_2}^*)\Im(\langle T^+\rangle )[(L^+_0S^0_0-L^+_sS^0_s)Y^0_0+(L^+_sS^0_sY^0_0)\Re(\zeta^\prime)]\nonumber\\
&&+2\Re(\Pi_\rho\Pi_{f_2}^*)[\Re(\langle V^+\rangle \langle T^+\rangle^*)-(L^+_sT^0_sY^0_2)\Re(\langle V^+\rangle )+(L^+_sT^0_sY^0_2)\Re(\langle V^+\rangle )\Re(\zeta^\prime)\nonumber\\
&& -(L^+_sV^0_sY^0_1)\Re(\langle T^+\rangle )\Re(1-\zeta^\prime)+(L^+_sV^0_sY^0_1)(L^+_sT^0_sY^0_2)|1-\zeta^\prime|^2]\nonumber\\
&&-2\Im(\Pi_\rho\Pi_{f_2}^*)[\Im(\langle V^+\rangle \langle T^+\rangle^*)-(L^+_sT^0_sY^0_2)\Im(\langle V^+\rangle )+(L^+_sT^0_sY^0_2)\Im(\langle V^+\rangle )\Re(\zeta^\prime)\nonumber\\
&& +(L^+_sV^0_sY^0_1)\Im(\langle T^+\rangle )\Re(1-\zeta^\prime)],\nonumber\\
\end{eqnarray}
and the CP-odd quantity ${\cal D}$ is
\begin{eqnarray}
{\cal D}=2\Im(\zeta^\prime)C(q^2,s_{_M})\Delta,
\end{eqnarray}
with
\begin{eqnarray}
\Delta&=&2\bigg[
\Im(\langle V^+\rangle )\{(L^+_sV^0_sY^0_1)|\Pi_\rho|^2+(L^+_sS^0_sY^0_0)\Re(\Pi_{f_0}\Pi_\rho^*)+(L^+_sT^0_sY^0_2)\Re(\Pi_\rho\Pi_{f_2}^*)\}\nonumber\\
&&+\Im(\langle T^+\rangle )\{(L^+_sT^0_sY^0_2)|\Pi_{f_2}|^2+(L^+_sS^0_sY^0_0)\Re(\Pi_{f_0}\Pi_{f_2}^*)+(L^+_sV^0_sY^0_1)\Re(\Pi_\rho\Pi_{f_2}^*)\}\nonumber\\
&&+\Re(\langle V^+\rangle )\{(L^+_sT^0_sY^0_2)\Im(\Pi_\rho\Pi_{f_2}^*)-(L^+_sS^0_sY^0_0)\Im(\Pi_{f_0}\Pi_\rho^*)\}\nonumber\\
&&-\Re(\langle T^+\rangle )\{(L^+_sV^0_sY^0_1)\Im(\Pi_\rho\Pi_{f_2}^*)+(L^+_sS^0_sY^0_0)\Im(\Pi_{f_0}\Pi_{f_2}^*)\}\nonumber\\
&&+(L^+_0S^0_0Y^0_0)(L^+_sV^0_sY^0_1)\Im(\Pi_{f_0}\Pi_\rho^*)+(L^+_0S^0_0Y^0_0)(L^+_sT^0_sY^0_2)\Im(\Pi_{f_0}\Pi_{f_2}^*)\bigg].
\end{eqnarray}
Note that since every term in $\Delta$ of Eq. (90) contains square terms of
$L^+_i$ which are proportional to $m_l$ (see Eq.~(\ref{Lpm})),
the CP-odd quantity ${\cal D}$ of Eq. (89)
is proportional to lepton mass
due to the definition of $\zeta^\prime$ (see Eq. (\ref{zetap})).
\end{appendix}
\newpage
|
\section{} command!!!
\newcommand{\sect}[1]{\setcounter{equation}{0}\section{#1}}
\renewcommand{\theequation}{\thesection.\arabic{equation}}
\newtheorem{predl}{Proposition}[section]
\newtheorem{defi}{Definition}[section]
\newtheorem{rem}{Remark}[section]
\newtheorem{cor}{Corollary}[section]
\newtheorem{lem}{Lemma}[section]
\newtheorem{theor}{Theorem}[section]
\vspace{0.3in}
\begin{flushright}
ITEP-TH-18/99\\
\end{flushright}
\vspace{10mm}
\begin{center}
{\Large\bf
Non-autonomous Hamiltonian systems related to highest Hitchin integrals}
\footnote
{Contribution in the Proceedings
"International Seminar on Integrable systems". In memoriam Mikail V. Saveliev.
Bonn, February, 1999}
\vspace{5mm}
A.M.Levin\\
{\sf Max-Plank-Institut f\"{u}r Matematik, Bonn}\\
{\sf Institute of Oceanology, Moscow, Russia,} \\
{\em e-mail <EMAIL>}\\
M.A.Olshanetsky
\\
{\sf Max-Plank-Institut f\"{u}r Matematik, Bonn}\\
{\sf Institute of Theoretical and Experimental Physics, Moscow, Russia,}\\
{\em e-mail <EMAIL>}\\
\vspace{5mm}
\end{center}
\begin{abstract}
We describe non-autonomous Hamiltonian systems coming from the Hitchin
integrable systems. The Hitchin integrals of motion depend on
the ${\cal W}$-structures of the basic curve. The parameters of the
${\cal W}$-structures play the role of times. In particular, the
quadratic integrals dependent on the complex structure (${\cal W}_2$-structure)
of the basic curve and times are coordinate on the Teichm\"{u}ller space.
The corresponding flows are the monodromy preserving equations such as
the Schlesinger equations, the Painlev\'{e} VI equation and their
generalizations. The equations corresponding to the highest integrals are
monodromy preserving conditions with respect to changing of the
${\cal W}_k$-structures ($k>2$). They are derived by the symplectic reduction
from the gauge field theory on the basic curve interacting with
${\cal W}_k$-gravity. As by product we obtain the classical Ward identities
in this theory.
\end{abstract}
\vspace{0.3in}
\begin{flushright}
{\it In memory of Mikhail Saveliev}\\
\end{flushright}
\vspace{0.12in}
\bigskip
\section {Introduction}
\setcounter{equation}{0}
Infinite-dimensional symmetries corresponding to the ${\cal W}$-algebras
play a central role in two-dimensional physics (see \cite{Hu} for a review).
Here we investigate classical Hamiltonian systems incorporated in a gauge
theory on a Riemann curve interacting with the ${\cal W}$ gravity.
Whereas the ${\cal W}_2$-gravity has a natural geometric background, there is
no satisfactory geometric understanding of the origin of the
${\cal W}_k,~(k>2)$ theory. There exists a viewpoint on the
${\cal W}_k$-gravity as the geometry of certain two-dimensional surfaces
embedded in $k$-dimensional affine space. One of Misha Saveliev's works
\cite{Sa} is an important step in this direction. Here we analyze
interrelations between ${\cal W}$-geometry and integrable systems.
On the later subject Misha Saveliev has had a significant influence.
We starting from some subclass of classical completely integrable system
with phase flows having the Lax form
\beq{1.1}
\partial_sL=[L,M_s],~~\partial_s=\frac{\partial}{\partial t_s}.
\end{equation}
We assume that \\
i)$L$ takes value in a complex Lie algebra. We restrict ourself to the
case ${\rm sl}(N,{\bf C})$;\\
ii)$L=L(z)$, where $z$ is a spectral parameter lying on the Riemann
curve $\Sigma_{g,n}$ of genus $g$ with $n$ marked points. $L$ can have first
order poles in the marked points. Thereby, we exclude in what follows the
Stokes phenomena
\footnote{This restriction was partly resolved in
the recent paper \cite{U}.}.
The integrable systems on curves without marked points was considered by
Hitchin \cite{H1}.
In this case nontrivial equations arise only on the high genus curves $(g>1)$.
Up to the recent time there were no explicit examples of such type
integrable systems\footnote{See, however, \cite{GP,Ga} for
$g=2,~L\in sl(2,{\bf C})$ }.
The generalization of Hitchin approach to the matrices with first order poles
in the marked points \cite{Ne} allowed to include in this scheme some well
known completely integrable models like the Toda and the Calogero-Moser systems.
The quantum counterpart of these systems are the Knizhnik-Zamolodchikov-Bernard
equations for conformal blocks in the WZW theory on the critical level,
while the original Hitchin systems correspond to equations for partition functions.
In the Hitchin type systems ${\rm tr} (L^k)=<L^k>$ being integrated over $\Sigma_{g,n}$
are integrals of motion. Before the integration one
should take into account that $L$ are $(1,0)$-forms on
$\Sigma_{g,n}$ in some fixed complex structure. To integrate $<L^k>$ one
should multiply it on $(-k+1,1)$-differentials
$\rho_{k,s}=\rho_{k,s}(z,\bar{z})\partial_z^{k-1}\otimes d\bar{z}$. The index $s$ arises
in the following way. The operator
$\rho_{k,s}(z,\bar{z})\partial_z^{k-1}$ is defined in local coordinates of the point
$(z,\bar{z})$. The fields $\rho_{k,s}(z,\bar{z})\partial_z^{k-1}$ that can be represented as
$\bar{\partial}$ derivative do not contribute in the Hamiltonians.
In this way $\rho_{k,s}$ can be chosen from $H^1(\Sigma_g, \Gamma^{k-1})$,
where $\Gamma$ are vector fields on $\Sigma_{g,n}$ vanishing in the marked points.
Then $s$ enumerates the basis in $ H^1(\Sigma_g, \Gamma^{k-1})$.
Represent the differentials as
\beq{1.3}
\rho_{k,s}=t_{k,s}\rho_{s,k}^0,
\end{equation}
where $\{\rho_{s,k}^0\}$ is a fixed basis in $ H^1(\Sigma_{g,n}, \Gamma^{k-1})$.
The integrals of motion take the form
\beq{1.1b}
H_{k,s}=\f1{k}\int_{\Sigma_g}<L^k>\rho_{k,s}^0,~~
(k=1,\ldots,N,~s=1,\ldots).
\end{equation}
The important class of Hamiltonian equations occurs when $t_{k,s}$
are considered as "times". In this case $H_{k,s}$ become the Hamiltonians of
non-autonomous systems, where times are
related to the deformations of the internal structure of the base
curve $\Sigma_{g,n}$.
It turns out that the phase flows of these systems are described by the
following modification of the Lax equation (\ref{1.1})
\beq{1.1a}
\partial_aL-\partial M_a+[M_a,L]=0,~~a=(k,s), \partial=\frac{\partial}{\partial z}.
\end{equation}
Consider first the quadratic Hamiltonians $H_{2,s}$. In this case
$\rho_{2,s}=\mu_s$ are the $(-1,1)$-differentials
(the Beltrami differentials).
They are $(0,1)$-forms taking values in the vector fields on $\Sigma_{g,n}$.
Here we deal with the Lie algebra of vector fields and
the group of local diffeomorphisms of $\Sigma_{g,n}$. Roughly speaking the space
$ H^1(\Sigma_{g,n}, \Gamma)$ can be defined as the space of smooth
$(-1,1)$-differentials on $\Sigma_{g,n}$ modulo global diffeomorphisms action.
The elements from $ H^1(\Sigma_{g,n}, \Gamma)$ play role of deformation
parameters of the complex structure on $\Sigma_{g,n}$.
In the genus zero case the complex structure is defined by the positions of
marked points $t_s=x_s-x_s^0$ and (\ref{1.1a}) leads to
the Schlesinger equation. Another interesting examples including a particular
family of the
Painlev\'{e} VI equation occur when the basic curve
is an elliptic curve with marked points \cite{LO}.
For $k>2$ the differentials $\rho_{k,s}$ do not
generate a Lie algebra. Due to this fact they have not natural geometric
description. Along with the dual objects (opers \cite{BD}) they generate the
so called ${\cal W}_k$-geometry of the basic curve $\Sigma_{g,n}$ \cite{F,GLM}.
${\cal W}_k$-geometry is a generalization of ${\cal W}_2$-geometry
that coincides with the space of projective structures of $\Sigma_{g,n}$.
The invariant object associated with the Lax form (\ref{1.1})
of integrable systems is the spectral curve ${\cal C}$
$$
{\cal C}: ~f(\lambda,z):=\det(\lambda+L(z))=0.
$$
It is $N$ cover of the basic curve $\Sigma_{g,n}$. Then the Prym variety
$Prym({\cal C}/\Sigma_g)$
is the Liouvillian torus of the completely integrable system (\ref{1.1})
\cite{H1}.
There are different parametrizations of ${\cal C}$. The two standard
parametrizations are the set
of the Hamiltonians $H_{k,s}$ (\ref{1.1b}) and the action variables.
The differentials $\rho_{k,s}$
(\ref{1.3}) provide another parametrization of ${\cal C}$ related to the
${\cal W}_N$ structure of the basic curve. For small times (\ref{1.1a})
describes an evolution of ${\cal C}_t$ near the fixed curve ${\cal C}_0$ .
The main goal of this paper is to investigate the dynamical systems
(\ref{1.1a}) associated
with the ${\cal W}_k$-geometry. There are two important aspects of this
investigation. First, the quantum analog of these systems are higher order
Knizhnik-Zamolodchikov-Bernard equations beyond the critical level.
It means that conformal
blocks in the WZW theory satisfy some analogue of nonstationar
Schr\'{o}dinger equations with higher order Casimirs and times
$t_{k,s},k>2$. We don't aware of some explicit examples of these type of
equations. On the other hand, the investigation of the higher
${\cal W}$-geometries
is interesting by itself. It is by no means an easy problem because apparently
these geometries are not of the Klein type - there are no evident group
symmetries related to them. The connection of the ${\cal W}$-geometries with the
integrable systems opens a new way for investigations of the
${\cal W}$-geometries. Namely, one can apply tools developed for integrable
systems, such as the
Whitham quantization method \cite{Kr1} which is well adjusted to
the analysis of the perturbations of integrals of motion
(\ref{1.1b}).
Our construction is based on the approach developed in \cite{F}.
The ${\cal W}_k$ structures are described there as a result of Hamiltonian
reduction with respect to a maximal parabolic subgroups of $\SL{k}$.
We generalize this approach in two directions:\\
$\bullet$ we consider Riemann curves with marked points
adding some additional data in the marked points;\\
$\bullet$ in addition to the ${\cal W}$ fields we include gauge fields.\\
As a result we obtain the classical Ward identities for ${\cal W}$-gravity
interacting with gauge theory on Riemann curves with marked points.
To obtain monodromy preserving flows we exclude the half
of ${\cal W}$ fields leaving only differentials $\rho_{k,s}$, which play
role of "times".
This procedure is also based on the Hamiltonian reduction with respect
to symmetries generated by the Sugawara type constraints.
On this stage the flows are rather trivial, since the systems are free.
Finally, the Hamiltonian reduction based on the gauge symmetries leads
to nontrivial dynamical systems.
The plan of the paper is follows. In the first sections we revise
the derivation of the equations preserving monodromies using the
projective structure of basic curves.
The similar program is done further in details for the ${\cal W}_3$ case.
In conclusion we discuss shortly the general case.
\section{Projective structures on Riemann curves and symplectic geometry}
\setcounter{equation}{0}
{\bf 1. Projective structures.}\\
Let us fix the complex structure on $\Sigma_{g,n}$ by choosing a pair
of local coordinates $(z,\bar{z})$ and the corresponding operators
$(\partial,\bar{\partial})$.
The deformed complex structure
can be read off from the solutions of the Beltrami equation
\beq{3.0}
(\bar{\partial}+\mu\partial)F=0.
\end{equation}
Locally, $F(z,\bar{z})$ is the diffeomorphism
$$
w=F(z,\bar{z}),~\bar{w}=\bar{F}(z,\bar{z})
$$
and the Beltrami differential is
$$
\mu(z,\bar{z})=-\frac{\bar{\partial}\bar{F}}{\partial F}.
$$
It defines the new complex structure operator $\partial_{\bar{w}}:=\bar{\partial}+\mu\partial$.
Two Beltrami differentials produce equivalent complex structures if they are
related by a global holomorphic diffeomorphisms of $\Sigma_{g,n}$
\beq{hd}
w=z-\epsilon(z,\bar{z}),~\bar {w}=\bar{z}.
\end{equation}
We assume that the Lie algebra ${\cal V}_{g,n}$ of corresponding vector fields
on $\Sigma_{g,n}$ is specified by their behavior near the marked points
\beq{3.01}
{\cal V}_{g,n}=\{\epsilon(z,\bar{z})\partial~|~\epsilon(z,\bar{z})=O(z-x_a)\}
\end{equation}
Under the holomorphic diffeomorphisms $\mu$ transforms as $(-1,1)$-differential.
The vector fields act on $\mu$ as
\beq{3.00}
j_{\epsilon}\mu=-\epsilon\partial\mu +\mu\partial\epsilon+\bar{\partial}\epsilon.
\end{equation}
We specify the dependence of $\mu$ on the positions of
the marked points in the following way.
Let ${\cal U}'_a\supset{\cal U}_a$ be two vicinities
of the marked point $x_a$
such that ${\cal U}'_a\cap{\cal U}'_b=\emptyset$ for $a\neq b$.
Let $\chi_a(z,\bar{z})$ be a smooth function
\beq{cf}
\chi_a(z,\bar{z})=\left\{
\begin{array}{cl}
1,&\mbox{$z\in{\cal U}_a$ }\\
0,&\mbox{$z\in\Sigma_g\setminus {\cal U}'_a.$}
\end{array}
\right.
\end{equation}
Introduce times related to the positions of the
marked points $t_{2,a}=x_a-x_a^0$.
Then $\mu$ can be represented as
\beq{6.1}
\mu=\sum_{a=1}^nt_{2,a}\mu^0_a,~~
\mu^0_a=\bar{\partial} \epsilon_a(z,\bar{z}),~~\epsilon_a(z,\bar{z})=\chi_a(z,\bar{z}),~~(t_{2,a}=x_a-x_a^0).
\end{equation}
Let $T$ be the
projective connection on $\Sigma_{g,n}$, i.e. $T$ is transformed under
the holomorphic diffeomorphisms as
$(2,0)$-differential up to the addition of the Schwarzian derivative.
Locally it means that
\beq{3.1b}
j_{\epsilon}T(z,\bar{z})=-\epsilon\partial T-2T\partial\epsilon-\frac{\kappa^2}{2}\partial^3\epsilon.
\end{equation}
Here $\kappa$ is a parameter, which later will play role of the "Planck constant"
in the Whitham quantization.
We assume that $T$ has poles at the marked points $x_a,(a=1,\ldots,n)$
up to the second order:
\beq{3.1a}
T|_{z\rightarrow x_a}\sim\frac{T_{-2,a}}{(z-x_a)^2}+\frac{T_{-1,a}}{(z-x_a)}+\ldots
\end{equation}
Let $\ti{\cal W}_2$ be the space of the pairs $(T,\mu)$ on $\Sigma_{g,n}$ with the
behavior near the marked points defined by (\ref{6.1}),(\ref{3.1a}).
\begin{defi}
The space ${\cal W}_2$ of projective structure on $\Sigma_{g,n}$ is the subset
of $\ti{\cal W}_2$ that satisfies the equation
\beq{3.1}
(\bar{\partial}+\mu\partial+2\partial\mu)T=\frac{1}{2}\partial^3\mu,
\end{equation}
with fixed values of
${\bf T}_2=(T_{-2,1},\ldots,T_{-2,l})$ in (\ref{3.1a}).
\end{defi}
Let $\psi$ be a $(-\frac{1}{2},0)$ differential. Then (\ref{3.1})
is the compatibility condition for the linear system
\beq{3.2}
(\bar{\partial}+\mu\partial -\frac{1}{2}\partial\mu) \psi=0,
\end{equation}
\beq{3.3}
(\partial^2-T)\psi=0.
\end{equation}
Consider two linear independent solutions $\psi_1,\psi_2$ to the
system. The projective structure $(T,\mu)$ can be equivalently defined
by their ratios $F=\psi_1/\psi_2$. In fact, it follows from (\ref{3.2}) that
$F$ satisfies the Beltrami equation (\ref{3.1}). Therefore, $\mu=\bar{\partial} F/\partial F$.
On the other hand, from (\ref{3.3}) $\psi_1=(\partial F)^\frac{1}{2}$ and $T={\cal S}_z(F)$,
where ${\cal S}_z(F)$ is the Schwarzian derivative of $F$. Arbitrary
linear independent solutions of the system (\ref{3.2}),(\ref{3.3}) are
obtained from $\psi_1,\psi_2$ by the $\SL{2}$ transform. It results in the
M\"{o}bius transform of $F$ and does not
change $T={\cal S}_z(F)$. Thus, the relations of any two independent solutions
defines projective structure as well.
\bigskip
\noindent
{\bf 2. Symplectic reduction with respect to diffeomorphisms.}\\
The space ${\cal W}_2$ is similar to the space of flat connections on $\Sigma_{g,n}$
for some gauge group. It is a symplectic manifold,
which can be derived from the affine space of smooth connections via
the symplectic reduction. The flatness condition plays the role
of the moment constraint equation. The similar procedure can be applied to
the space $\ti{\cal W}_2$ to obtain ${\cal W}_2$. In this case the gauge group is
replaced by the group
of holomorphic diffeomorphisms of $\Sigma_{g,n}$ (\ref{hd}). The space
$\ti{\cal W}_2$ can be endowed with the symplectic structure
\beq{3.4}
\omega=-\kappa^{-1}\int_{\Sigma_g}\delta T\delta\mu.
\end{equation}
The action of the vector fields ${\cal V}_{g,n}$ (\ref{3.00}),(\ref{3.1b})
is the symmetry of $\omega$. The Hamiltonian of this action
$$
H_\epsilon=-\kappa^{-1}\int_{\Sigma_g}\epsilon
[(\bar{\partial}+\mu\partial+2\partial\mu)T-\frac{\kappa^2}{2}\partial^3\mu]
$$
produces the moment map
\beq{3.04}
m:\ti{W}_2\rightarrow {\cal V}^*_{g,n},~~m=(\bar{\partial}+\mu\partial+2\partial\mu)T-\frac{\kappa^2}{2}\partial^3\mu,
\end{equation}
where ${\cal V}^*_{g,n}$ is the dual space to the algebra ${\cal V}_{g,n}$
of vector fields.
It is the space of $(2,1)$-forms on $\Sigma_{g,n}$. As it follows from
(\ref{3.01}) in the neighborhoods of marked points elements $y\in{\cal V}^*_{g,n}$
take the form
\beq{3.03}
y\sim b_{1,a}\partial\delta(x_a)+b_{2,a}\partial^2\delta(x_a)+\ldots.
\end{equation}
Thereby, the terms $T_{-1,a}\delta(x_a)$ that arise in $\bar{\partial} T$ (\ref{3.04})
from the first order
poles of $T$ (\ref{3.1a}) are projected out from the moment map $m$ (\ref{3.04}).
We take
$$
m=-\sum_{a=1}^nT_{-2,a}\partial\delta(x_a).
$$
It follows from (\ref{3.1a}),(\ref{3.04}) and (\ref{3.03}) that we put in (\ref{3.03})
$ b_{1,a}=-T_{-2,a},~~b_{k,a}=0,~k>1.$
Thus, the condition (\ref{3.1}) that distinguish ${\cal W}_2$ in $\ti{\cal W}_2$
is the moment constraint with respect to the action of the diffeomorphisms.
\bigskip
\noindent
{\bf 3."Drinfeld-Sokolov" approach.}
In \cite{F} another procedure was proposed, which resembles
the Drinfeld-Sokolov approach.
Shortly, it looks as follows. Consider the affine space ${\cal N}_2$ of
$\SL2$ smooth flat connections on $\Sigma_{g,n}$
$$
{\cal N}_2=\{adz+\bar{a}d\bar{z}\},
$$
\beq{3.7}
F(a,\bar{a})=\bar{\partial} a-\partial\bar{a}+[a,\bar{a}]=0.
\end{equation}
The field $a$ has poles in the marked points up to the second order.
The space ${\cal N}_2$ has the standard symplectic form
\beq{3.5}
\omega'=\int_{\Sigma_g}{\rm tr}(\delta a\delta\bar{a}).
\end{equation}
The form is invariant under the gauge transform
$$
a\rightarrow g^{-1}ag+g^{-1}\partial g,~~\bar{a}\rightarrow g^{-1}\bar{a}g+g^{-1}\bar{\partial} g,~~
g\in{\cal G}={\rm Map}(\Sigma_{g,n},\SL2)
$$
We assume that the Lie algebra of the gauge group Lie$({\cal G})$ is
specialized by the behavior of its matrix element $x^{12}$ near the
marked points
\beq{3.5a}
x^{12}|_{z\rightarrow x_a}=O(z-x_a),~~(x^{JK})\in{\rm Lie}({\cal G}).
\end{equation}
The flatness (\ref{3.7}) is the moment constraint with respect to the action
of the gauge group.
The form $\omega'$ is degenerated - it vanishes on the orbits
of the gauge
group, because we only put the moment condition (\ref{3.7}) and do not
fix the gauge. If we do it we come to the finite-dimensional space of flat
connections, but it is necessary to leave two fields $T$ and $\mu$.
The trick proposed in \cite{F} is to fix the gauge with respect to
the Borel subgroup $B$ of the lower triangular matrices.
It was proved there that it allows to obtain from the space ${\cal N}_2$
the space of projective connections ${\cal W}_2$.
The gauge freedom allows to fix a generic matrix $a$ in the form
\beq{3.6}
a=\mat{0}{1}{T}{0},
\end{equation}
where $T$ is a new field satisfying (\ref{3.1a}). The first order poles of
$T$ do not contribute in the moment equation (\ref{3.7}) since they are eaten
by the gauge transform (see (\ref{3.5a})).
The moment equation (\ref{3.7})
allows to express all matrix elements of $\bar{a}$ in terms of $T$
and new field $\mu$
\beq{3.8}
\bar{a}=\mat{\frac{1}{2}\partial\mu}{-\kappa^{-1}\mu}{-\kappa^{-1}\mu T+\frac{1}{2}\kappa\partial^2\mu}{-\frac{1}{2}\partial\mu}.
\end{equation}
The flatness condition becomes trivial for all matrix element except
$F_{(2,1)}$. It can be checked that it just coincides with the projectivity
condition (\ref{3.1}).
The linear system for $(-\frac{1}{2},0)$ differentials
$\psi$
\beq{3.9a}
(\bar{\partial}+\bar{a})
\left(\begin{array}{c}
\psi\\
\partial \psi
\end{array}\right)=0,
~~(\kappa\partial+a)
\left(\begin{array}{c}
\psi\\
\partial \psi
\end{array}\right)=0
\end{equation}
is consistent due to (\ref{3.7}). It is the matrix form
of (\ref{3.2}),(\ref{3.3}) for the special form
of $a$ and $\bar{a}$ (\ref{3.6}),(\ref{3.8}).
The original symplectic structure $\omega'$
(\ref{3.5}) is reduced to $\omega$ (\ref{3.4}) on the space of the projective
structures ${\cal W}_2$.
Though, the diffeomorphisms do not arise in this approach, they are hidden
in this construction. To demonstrate it, calculate the commutator
of two matrices $\bar{a}_1$, $\bar{a}_2$
$$
[\bar{a}_1(\mu_1,T_1),\bar{a}_2(\mu_2,T_2)]_{(1,2)}=
\mu_1\partial\mu_2-\mu_2\partial\mu_1.
$$
Thus, the commutator of matrix $\bar{a}$ reproduces the commutator of
vector fields.
\section {Isomonodromic deformations and projective structures}
\setcounter{equation}{0}
{\bf 1.${\cal W}_2^N$-structures - definition.}\\
Consider some projective structure on $\Sigma_g$ defined by the linear system
(\ref{3.2}),(\ref{3.3}). We generalize it in the following way.
Consider the vector $\SL{N}$-bundle $V$
over $\Sigma_g$.
Let $(A,\bar{A})$ be connections in $V$ corresponding to the complex structure
we have fixed.\\
We assume that\\
$\bullet$ the connection $\bar{A}$ is smooth;\\
$\bullet$ the connection $A$ has first order poles in the marked points
\beq{4.0}
A\sim\frac{A_{-1,a}}{z-x_a}+A_{0,a}+\ldots,~a=1,\ldots,n.
\end{equation}
In addition to this data consider the set of coadjoint orbits of $\SL{N}$
in the marked points
\beq{4.00}
({\cal O}_1,\ldots,{\cal O}_n),~~
{\cal O}_a=\{(p_a=g_ap_a^0g_a^{-1}|g_a\in\SL{N}\}.
\end{equation}
Here $p_a^0$ specifies the choice of the orbit ${\cal O}_a$.
To reconcile this data with the projective structures instead of the operator
(\ref{3.3})
we consider in what follows the matrix operator
$$
(\kappa\partial+A)^2-T.
$$
Define the matrices
\beq{4.3}
\ti{T}=T-A^2-\kappa\partial A.
\end{equation}
and
\beq{f11}
f_1=-\bar{A}+\frac{1}{2}\partial\mu{\bf 1}_N-\f1{\kappa}\mu A.
\end{equation}
We skip the multiplication on the scalar matrices ${\bf 1}_N$ in what follows.
Define the space $\ti{\cal W}_2^N$ of fields
$(T,\mu),(A,\bar{A}),{\bf p}=(p_1,\ldots,p_l)$
on $\Sigma_{g,n}$, where the behavior of $\mu,T$ and $A$ in the neighborhood
of marked points satisfies (\ref{6.1}),(\ref{3.1a}) and (\ref{4.0})
correspondingly.
\begin{defi}
${\cal W}_2^N$-structure on $\Sigma_g$ is the subset of $\ti{\cal W}_2^N$
satisfying the following identities:
\beq{4.5}
(\bar{\partial}+\mu\partial+2\partial\mu)\ti{T}-\frac{1}{2}\kappa^2\partial^3\mu=[\ti{T},\bar{A}+\f1{\kappa}\mu A]
+2\kappa A\partial f_1 ,
\end{equation}
\beq{4.6}
\bar{\partial} A-\kappa\partial\bar{A}+[A,\bar{A}]=0,
\end{equation}
\beq{4.6a}
A_{-1,a}=-p_a,~(a=1,\ldots,n),
\end{equation}
\beq{4.6b}
T_{-2,a}=\frac{<(p_a^0)^2>}{N}
~(a=1,\ldots,n).
\end{equation}
\end{defi}
In the first relation the right hand side is the contribution of the
gauge fields in the classical Ward identity for the gravitational fields
$(T,\mu)$. The second identity is the standard flatness condition for
the gauge fields. The last identity expresses the most singular terms
$T_{-2,a}$ of the projective connection $T$ in terms of the second Casimirs
parametrizing the orbits ${\cal O}_a$.
There is a straightforward generalization of the linear system
(\ref{3.2}),(\ref{3.3}) defining the projective structure.
Let $\Psi$ be the element of the space of sections
$\Omega^{(-\frac{1}{2},0)}(\Sigma_g,{\rm Aut}V)$.
\begin{predl}
The equations (\ref{4.5}),(\ref{4.6}),
are the compatibility conditions for the linear system
\beq{4.1}
(\bar{\partial}+\mu\partial -\frac{1}{2}\partial\mu+\bar{A}+\f1{\kappa}\mu A) \Psi=0,
\end{equation}
\beq{4.2}
[(\kappa\partial+A)^2-T]\Psi=0.
\end{equation}
\end{predl}
The space ${\cal W}_2^N$ can be endowed with the degenerated symplectic
form
\beq{4.7}
\omega=\int_{\Sigma_{g,n}}[-\frac{N}{\kappa}\delta T\delta\mu+
2<\delta A,\delta\bar{A}>]+4\pi i\sum_{a=1}^n\delta <p_a,g_a^{-1}\delta g_a>.
\end{equation}
The last sum in (\ref{4.7}) is the contribution of symplectic forms on
the coadjoint orbits. We demonstrate below that this form is natural and come
from a Hamiltonian reduction procedure.
\bigskip
\noindent
{\bf 2. Symplectic construction.}\\
Consider the vector bundle $V_{2N}=V_2\otimes V_N$
over $\Sigma_g$
with the structure group\\
$G=S(\GL{2}\otimes\GL{N})\sim \SL{2N^2}$. In the
nondeformed complex structure on $\Sigma_{g,n}$ the connection operators are
$$
\kappa\partial-{\cal A},~~\bar{\partial}-\bar{\cal A},
$$
where ${\cal A},\bar{\cal A}$ take values in Lie$(S(\GL{2}\otimes\GL{N}))$.
The first components of ${\cal A}$ and $\bar{\cal A}$ act on the sections of
$1$-jets of $\Omega^{(-\frac{1}{2},0)}(\Sigma_{g,n})$, while the second components act
on the sections of the bundle $\Omega^{(-\frac{1}{2},0)}(\Sigma_{g,n},{\rm Aut}V)$.
Define the space ${\cal K}_2$ as the space of connections
$({\cal A},\bar{\cal A})$
with coadjoint orbits attached in the marked points
$$
({\cal A},\bar{\cal A});
({\bf 1}_2\otimes{\bf p})=({\bf 1}_2\otimes p_1,\ldots,{\bf 1}_2\otimes p_l)
$$
with the following additional restrictions:\\
$\bullet$ the ${\cal A}$ component are gauge equivalent
to the special form
\beq{4.60a}
{\cal A}\sim DS\otimes{\bf 1}_N+{\bf 1}_2\otimes(-A),
\end{equation}
where
$$
DS=\mat{0}{1}{T}{0};
$$
$\bullet$ near the marked points $T$ has the form (\ref{3.1a});\\
$\bullet$ near the marked points $A$ has the form (\ref{4.00}).
Propositions 3.1 follow from the following statement
\begin{predl}
${\cal W}_2^N$ is the subset ${\cal K}_2$ satisfying the flatness condition
\beq{4.50}
[\kappa\partial-{\cal A},\bar{\partial}-\bar{\cal A}]=0.
\end{equation}
\end{predl}
{\sl Proof.}\\
The space ${\cal K}_2$ has the standard symplectic form
\beq{4.30}
\omega'=\int_{\Sigma_{g,n}}<<\delta{\cal A},\delta\bar{\cal A}>>+
4\pi i\sum_{a=1}^n\delta <p_a,g_a^{-1}\delta g_a>
\end{equation}
Here $<<,>>$ is the trace in the tensor product, and $<,>$ is the trace
in the $V_N$ space.
Introduce the following group of gauge transforms. It is the smooth maps
$$
{\cal G}=\{{\rm Map}(\Sigma_{g,n}, \SL{2N^2})\}
$$
with additional restrictions for the maps near the marked points. We formulate
them on the Lie algebra Lie$({\cal G})$ level. Let $x^{IJ}_{\alpha\beta}$ be
the matrix element
in Lie$({\cal G})$. Here the upper indices $I,J$ are related to the space $V_2$
and the lower indices $\alpha,\beta$ to the space $V_N$. We assume that in the
neighborhood of the marked point
\beq{ge2}
x^{12}_{\alpha\beta}\sim\delta_{\alpha\beta}O(z-x_a)+(1-\delta_{\alpha\beta})O(z-x_a)^2,
\end{equation}
while the other matrix element are continuae in the marked points.
It can be checked that these matrices define the Lie subalgebra
Lie$({\cal G})$ in the Lie algebra of smooth maps.
The form $\omega'$ (\ref{4.30}) is invariant under the gauge transform
$$
\bar{\cal A}\rightarrow f^{-1}\bar{\partial} f+ f^{-1}\bar{\cal A}f,
~~f\in{\cal G},
$$
\beq{gt}
{\cal A}\rightarrow f^{-1}\kappa\partial f+ f^{-1}{\cal A}f
\end{equation}
$$
g_a\rightarrow g_af_a,~p_a\rightarrow f_a^{-1}p_af_a,~f_a=f(x_a,\bar{x}_a).
$$
We identify the dual space $($Lie$)^*({\cal G})$ with matrices living
on $\Sigma_{g,n}$ by means of the bilinear form
$$
\int_{\Sigma_g}<<x,y>>,~(x\in{\rm Lie}({\cal G}),~y\in{\rm Lie}^*({\cal G}).
$$
Due to (\ref{ge2}) the matrix elements $y\in($Lie$)^*({\cal G})$ have
the following form in the marked points
\beq{ds2}
y^{21}_{\alpha\al}\sim\sum_{k\geq 1}b_k\partial^k \delta(x_a),~~
y^{21}_{\alpha\beta}\sim\sum_{k\geq 2}b_{k,\alpha,\beta}\partial^k \delta(x_a),~(\alpha\neq\beta).
\end{equation}
\beq{dds2}
y^{JK}_{\alpha\beta}\sim\sum_{k\geq 0}b_{k,\alpha,\beta}\partial^k \delta(x_a),
~~(J\neq 2,~K\neq 1).
\end{equation}
The form $\omega'$ (\ref{4.30}) is degenerated on the orbits of the gauge group.
The condition (\ref{4.60a}) is similar to the partial gauge fixing.
We choose ${\cal A}$ in the form
\beq{4.8}
{\cal A}=\mat{0}{E}{\ti{T}}{-2A},
\end{equation}
where $\ti{T}$ is given by (\ref{4.3}). This form of ${\cal A}$ is the gauge
transform of (\ref{4.60a}) by
$$
f=\mat{{\bf 1}_N}{0}{A}{{\bf 1}_N}.
$$
As usual, the curvature tensor $F({\cal A},\bar{\cal A})$ is the moment map
$$
m:({\cal A},\bar{\cal A},{\bf p})\rightarrow({\rm Lie})^*({\cal G}).
$$
The flatness (\ref{4.50}) means that we take $m=0$.
It allows to define $\bar{\cal A}$.
The general solution of (\ref{4.50}) depends on two fields $\bar{A}$ and $\mu$
and takes the form
\beq{4.10}
\bar{\cal A} =\mat{-\bar{A}+\frac{1}{2}\partial\mu-\kappa^{-1}\mu A}{-\kappa^{-1}\mu}
{-\kappa^{-1}\mu\ti{T}-\kappa\partial(\bar{A}-\frac{1}{2}\partial\mu+\kappa^{-1}\mu A)}
{-\bar{A}+\kappa^{-1}\mu A-\frac{1}{2}\partial\mu},
\end{equation}
where $\bar{A}$ takes value in Lie$(\SL{N})$.
It can be checked immediately that (\ref{4.50}) becomes the identity for
the $\GL{2}$ matrix elements $(1,1)$ and $(1,2)$ and coincides with
(\ref{4.5}) and (\ref{4.6}) for $(2,1)$ and $(2,2)$ respectively.
Now let us discuss the boundary condition (\ref{4.6a}),(\ref{4.6b}).
Since $m=0$ all coefficients in
(\ref{ds2}) and (\ref{dds2}) of
$F^{2,1}({\cal A},\bar{\cal A})$ and $F^{2,2}({\cal A},\bar{\cal A})$
vanish. The expression $\bar{\partial} \ti{T}$ in the matrix element $(2,1)$
has the terms proportional to $\partial \delta(x_a)$. Their cancellation lead
to (\ref{4.6b}). In the similar way, using (\ref{dds2})
we obtain (\ref{4.6a}) from $\bar{\partial} A$ in the matrix element $(2,2)$.$\Box$
The flatness of these connections is equivalent to the compatibility of the
linear system
\beq{4.9a}
(\kappa\partial-{\cal A})\left(
\begin{array}{c}
\psi \\
\kappa\partial\psi
\end{array}
\right )=0,
\end{equation}
\beq{4.9b}
(\bar{\partial}-\bar{\cal A})\left(
\begin{array}{c}
\psi \\
\kappa\partial\psi
\end{array}
\right )=0
\end{equation}
Substituting in this system ${\cal A}$ (\ref{4.8}) and $\bar{\cal A}$
(\ref{4.10}) we obtain the linear system (\ref{4.1}), (\ref{4.2})
in Proposition 3.1.
The reduced symplectic form is read off from (\ref{4.30}),(\ref{4.8})
and (\ref{4.10}). It coincides with $\omega$ (\ref{4.7}).
\bigskip
\noindent
{\bf 3. Isomonodromic deformations.}\\
To obtain nontrivial dynamical systems with Hamiltonians
we modified our previous procedure. Instead of the
gauge group ${\cal G}$ defined in (\ref{gt}) consider its parabolic subgroup
$$
{\cal G}^P=\left\{\Sigma_{g,n}\rightarrow \mat{a}{0}{b}{a^{-1}}\otimes\SL{N}\right\},
$$
Due to this choice
there are no moment constraints coming from the matrix elements $(2,1)$ of
$\GL{2}$.
Therefore, the conditions (\ref{4.5}),(\ref{4.6b}) are absent in this case.
In this way we come to the manifold ${\cal K}_2^N$ with
the constraints
(\ref{4.6}),(\ref{4.6a}) and the symplectic form (\ref{4.7}). Evidently
${\cal W}_2^N\subset{\cal K}_2^N\subset{\cal K}_2$.
Consider the following transformations of the fields in ${\cal K}_2^N$
\beq{4.11}
\mu\rightarrow \mu+\xi,~~\bar{A}\rightarrow\bar{A}+\f1{\kappa}A\xi,~~T\rightarrow T,~~ A\rightarrow A,
\end{equation}
where $\xi\in\Omega^{(-1,1)}(\Sigma_g)$.
It is the symmetry of the form $\omega$ (\ref{4.7}).
The transformations (\ref{4.11}) are generated by the Hamiltonian
$$
{\cal H}_\xi=\int_{\Sigma_{g,n}}\xi(T-\frac{<A^2>}{N}).
$$
We put ${\cal H}_\xi=0$. It means, that the Sugawara relation
\beq{4.12a}
T=\f1{N}<A^2>
\end{equation}
is the moment constraint with respect to the symmetry (\ref{4.11}).
We don't fix the gauge and thereby come to the space with fields
$$
{\cal P}^N_2=\{A,\bar{A},{\bf p},\mu\}\sim{\cal K}_2^N/(T=\f1{N}<A^2>) .
$$
It is a bundle over the space of times $\{\mu\}$.
It follows from (\ref{4.7}) that on this space $\omega$ take the form
\beq{4.13}
\frac{1}{2}\omega=\int_{\Sigma_g}<\delta A,\delta\bar{A}>
+2\pi i\sum_{a=1}^n\delta<p_a,g_a^{-1}\delta g_a>
-\f1{2\kappa}\int_{\Sigma_g}\delta<A^2>\delta\mu.
\end{equation}
Since we
did not fix the gauge of this transformation and preserve the field
$\mu$ the form $\omega$ (\ref{4.13}) is degenerated on ${\cal P}^N_2$.
If one replaces
\beq{4.14a}
\bar{A}'=\bar{A}-\f1{\kappa}\mu A,
\end{equation}
(\ref{4.13}) takes the canonical form
\beq{4.13b}
\frac{1}{2}\omega=\int_{\Sigma_g}<\delta A,\delta\bar{A}'>+2\pi i\sum_{a=1}^n\delta<p_a,g_a^{-1}\delta g_a>.
\end{equation}
Now we pass to the finite-dimensional space of equivalent
complex structures
${\cal T}_{g,n}$ (the Teichm\"{u}ller space).
The tangent space to the Teichm\"{u}ller space is
$H^1(\Sigma_g, \Gamma)$. Note, that only elements
of $H^1(\Sigma_g, \Gamma)$ contribute in the second integral
in (\ref{4.13}). According to the
Riemann-Roch theorem $H^1(\Sigma_g, \Gamma)$ has dimension
\beq{dim2}
l_2=\dim H^1(\Sigma_g, \Gamma)=3g-3+n.
\end{equation}
We fix a reference point
${\bf \mu}^0=(\mu^0_1,\ldots,\mu^0_{l_2})$.
Then
\beq{cst}
\mu=\sum_{s=1}^{l_2}t_s\mu_s^0.
\end{equation}
defines a new complex structure and ${\bf t}=(t_1,\ldots,t_{{l_2}})$
are coordinates of the tangent vector to ${\bf \mu}^0$ in
$H^1(\Sigma_g, \Gamma)$.
Expanding $\mu$ in the basis (\ref{cst}) we rewrite $\omega$ as
\beq{4.14}
\frac{1}{2}\omega=\omega^0-\f1{\kappa}\sum_s\delta H_s\delta t_s,~~
H_s=H_s(A,{\bf t})=\frac{1}{2}\int_{\Sigma_g}<A^2>\mu_s^0,
\end{equation}
where
$$
\omega^0=\int_{\Sigma_g}<\delta A,\delta\bar{A}>+2\pi i\sum_{a=1}^n\delta<p_a,g_a^{-1}\delta g_a>,~~
({\bf t}=(t_1,\ldots,t_l),~\partial_s=\partial/\partial t_s).
$$
We keep the same notations for
the space ${\cal P}^N_2$
$$
\begin{array}{cc}
{\cal P}^N_2& \\
\downarrow& \sim{\cal R}=\{A,\bar{A},{\bf p}\}\\
{\cal T}_g &
\end{array}
$$
${\cal P}^N_2$ is the extended phase space.
The form $\omega^0$ is nondegenerated on the fibers. In this situation
the equation of motion for any function $F$ on ${\cal P}^N_2$ takes the form
\cite{Ar}
$$
\kappa\frac{d F}{d t_s} =\kappa\frac{\partial F}{\partial t_s}+\{H_s,F\}_{\omega^0}.
$$
In addition, there are the consistency conditions for the Hamiltonians
({\em the Whitham equations})
\beq{WE}
\kappa\partial_sH_r-\kappa\partial_rH_s+\{H_r,H_s\}=0.
\end{equation}
They are satisfied since the Hamiltonians (\ref{4.14}) commute.
On this stage the equations of motion are trivial.
\beq{4.18}
\partial_sA=0,~~\partial_s\bar{A}=\frac{1}{2} A\mu_s^0~~(s=1,\ldots,l).
\end{equation}
We call this system the {\em Hierarchy of the Isomonodromic Deformations}
(HID). This notion will be justified below.
The degenerated symplectic form (\ref{4.14}) is invariant under the
gauge transformations \\
$\ti{\cal G}={\rm Map}(\Sigma_{g,n}\SL{N})$
\beq{4.17}
A\rightarrow f^{-1}Af+f^{-1}\kappa\partial f,~\bar{A}\rightarrow f^{-1}\bar{A} f+f^{-1}(\bar{\partial}+\mu\partial) f,~
p_a\rightarrow f^{-1}(x_a)p_af(x_a),~g_a\rightarrow g_af(x_a).
\end{equation}
The flatness condition
\beq{4.13a}
(\bar{\partial}+\partial\mu)A-\kappa\partial\bar{A}+[\bar{A},A]=0.
\end{equation}
is the moment constraint generating this symmetry.
It allows to consider the linear consistent system on $\Sigma_g$
\beq{4.19}
(\kappa\partial+A)\Psi=0,
\end{equation}
\beq{4.20}
(\bar{\partial}+\sum_st_s\mu^0_s\partial+\bar{A})\Psi=0,
\end{equation}
where $\Psi\in\Omega^0(\Sigma_{g,n},V)$.
The monodromy matrix $Y\in{\rm Rep}(\pi_1(\Sigma_{g,n}))\to\SL{N}$ transforms
solutions as
$$
\Psi\rightarrow\Psi Y.
$$
The isomonodromy of (\ref{4.19}),(\ref{4.20}) is the independence of $Y$
on the deformations of the complex structure of $\Sigma_{g,n}$.
\begin{predl}
The linear system (\ref{4.19}),(\ref{4.20}) has the property of
the isomonodromic deformations with respect to the "times" $t_s$ iff
the equations of motion (\ref{4.18}) are satisfied. In other words,
(\ref{4.18}) are monodromy preserving equations.
\end{predl}
{\sl Proof}\\
It follows from (\ref{4.18}) that $\partial_s$ commute with $(\kappa\partial+A)$ and
$(\bar{\partial}+\sum_st_s\mu^0_s\partial+\bar{A})$. Thus, in addition to (\ref{4.19}),(\ref{4.20})
one has consistent equations
\beq{4.21}
\partial_s\Psi=0,~~(s=1,\ldots,l).
\end{equation}
Then it follows from (\ref{4.21}) that
\beq{4.22}
\partial_sY=0,~~(s=1,\ldots,l).
\end{equation}
Assume now that the monodromy of the linear system (\ref{4.19}),(\ref{4.20})
is the time independent (\ref{4.22}). Then (\ref{4.21}) is fulfilled. The first
equation of motion (\ref{4.18}) follows from (\ref{4.19}) and (\ref{4.21}) and
the second from (\ref{4.20}) and (\ref{4.21}). $\Box$
The equations (\ref{4.18}) are not very interesting, because they describe a free
motion. They become nontrivial after the Hamiltonian reduction with respect to
the gauge transformations (\ref{4.17}). The flatness (\ref{4.13}) is the moment
equation. Consider the gauge fixing assuming that $\bar L$ parametrizes the orbits
of the gauge group $\ti{\cal G}$ in the phase space ${\cal R}$
\beq{GF}
\bar{L}=f^{-1}\bar{A} f+f^{-1}(\bar{\partial}+\mu\partial) f.
\end{equation}
It allows partly to fix $f$.
At the same time the gauge transformation defines
$$
L=f^{-1}Af+f^{-1}\kappa\partial f.
$$
We keep the same notations for the transformed matrices $p_a$.
Substituting these two expressions in the moment equation we obtain
\beq{4.33}
(\bar{\partial}+\partial\mu)L-\kappa\partial\bar{L}+[\bar{L},L]=0,
\end{equation}
where according to (\ref{4.0}),(\ref{4.6a}) Res$L|_{z=x_a}=p_a$.
The gauge fixing (\ref{GF}) and the moment constraint (\ref{4.33})
kill almost all degrees of freedom. The fibers
${\cal R}^{red}=\{L,\bar{L},{\bf p}\}$ become finite-dimensional,
as well as the bundle
${\cal P}^{red,N}_2$.
The form $\omega$ (\ref{4.14}) on ${\cal P}^{red,N}_2$ is
\beq{red}
\omega^0=\int_{\Sigma_{g,n}}<\delta L,\delta\bar{L}>+
2\pi i\sum_{a=1}^n\delta<p_a,g_a^{-1}\delta g_a>,~~
H_s=H_s(L,{\bf t})=\frac{1}{2}\int_{\Sigma_{g,n}}<L^2>\mu_s^0.
\end{equation}
But now the system is no long free because
due to (\ref{4.33}) $L$ depends on $\bar{L}$ and ${\bf p}$.
The equations of motion (\ref{4.18}) take the form
\beq{4.34}
\kappa\partial_sL-\kappa\partial M_s+[M_s,L]=0,~~M_s=\partial_s ff^{-1},
\end{equation}
\beq{4.35}
\kappa\partial_s\bar{L}-(\bar{\partial}+\mu\partial)M_s+[M_s,\bar{L}]=L\mu^0_s.
\end{equation}
The equation (\ref{4.34}) is the analog of the Lax equation. The essential
difference is the differentiation with respect to the spectral parameter
$\partial$. The last equation determines the matrices $M_s$.
These equations are nontrivial and for the genus $g=0,1$ reproduce
the Schlesinger system, Elliptic Schlesinger system, multicomponent
generalization of the Painlev\'{e} VI equation \cite{LO}.
The equations (\ref{4.34}), (\ref{4.35}) along
with (\ref{4.33}) are consistency conditions for the linear system
\beq{4.36}
(\kappa\partial+L)\Psi=0,
\end{equation}
\beq{4.37}
(\bar{\partial}+\sum_st_s\mu_s^0\partial+\bar{L})\Psi=0,
\end{equation}
\beq{4.38}
(\partial_s+M_s)\Psi=0,~~(s=1,\ldots,l_2).
\end{equation}
As in Proposition 3.3 the equations (\ref{4.38}) provides the isomonodromy
property of the system (\ref{4.36}), (\ref{4.37}) with respect to variations
of the times $t_s$.
\bigskip
\noindent
{\bf 4.Scaling limit.}\\
Consider the limit $\kappa\to 0$. The value $\kappa=0$ is called critical.
The symplectic form $\omega$ (\ref{4.14}) is singular in this limit.
Let us replace the times
$$
t_s\rightarrow t_s^0+\kappa t_s,
$$
and assume that the times $t_s^0, ~(s=1,\dots)$ are fixed.
After this rescaling the form (\ref{4.14}) become regular.
The rescaling procedure means that we blow up a vicinity
of the fixed point $\mu_s^{(0)}$
in ${\cal T}_{g,n}$ and the whole dynamic
is developed in this vicinity.
This fixed point is defined by the complex coordinates
\beq{fp}
w_0=z-\sum_st^0_s\epsilon_s(z,\bar{z}),~~\bar{w}_0=\bar{z},~~
(\partial_{\bar{w}_0}=\bar{\partial}+\sum_st_s^0\mu_s^0).
\end{equation}
For $\kappa=0$ the connection $A$ is transformed into the one-form $\Phi$
(the Higgs field)
$
\kappa\partial +A\rightarrow \Phi,
$
(see (\ref{4.17})).
Let
$
L^0=\lim_{\kappa\to 0}L,~~\bar{L}^0=\lim_{\kappa\to 0}\bar{L}.
$
Then we obtain the autonomous Hamiltonian systems with the form
$$
\omega^0=\int_{\Sigma_g}<\delta L^0,\delta\bar{L}^0>+
2\pi i\sum_{a=1}^n\delta<p_a,g_a^{-1}\delta g_a>
$$
and the commuting quadratic integrals (\ref{WE}). The phase space
${\cal R}^{red}$ is the cotangent bundle to the
moduli of stable holomorphic $\SL{N}$-bundles over $\Sigma_{g,n}$. These
systems are completely integrable for the $\SL{2}$ bundles \cite{Ne}.
The corresponding set of linear equations has the following form.
The level $\kappa$ can be considered as the Planck constant (see
(\ref{4.36})). We consider the quasi-classical regime
$$
\Psi=\phi\exp\frac{\cal S}{\kappa},
$$
where $\phi$ is a group-valued function and ${\cal S}$ is a scalar phase.
Assume that
$$
\frac{\partial}{\partial\bar w_0}{\cal S}=0,~~
\frac{\partial}{\partial t_s}{\cal S}=0.
$$
In the quasi-classical limit we set
$$
\partial{\cal S}=\lambda.
$$
Then instead of (\ref{4.36}), (\ref{4.37}), (\ref{4.38}) we obtain
$$
(\lambda+L^0)\Psi=0,
$$
$$
(\bar{\partial}_{{\bar w}_0}+\lambda\sum_st_s\mu_s^0+\bar{L}^0)\Psi=0,
$$
$$
(\partial_s+M^0_s)Y=0,~~(s=1,\ldots,l_2).
$$
Note, that the consistency conditions for the first and the last equations
are the standard Lax equations (\ref{1.1}).
\section{${\cal W}_3$ and Isomonodromic deformations}
\setcounter{equation}{0}
\noindent
{\bf 1. ${\cal W}_3$-structure.}\\
Define the space $\ti{\cal W}_3$
of fields $(W,\rho),(T,\mu)$ on $\Sigma_{g,n}$,
where $(T,\mu)$ are the same as in the case of the projective structures,
$\rho$ is the $(-2,1)$-differential and a variation of $W$ is
$(3,0)$-differential \cite{F,GLM}.
We assume that behavior of $\rho$ near the marked points
has the form (compare with (\ref{6.1}))
\beq{5.00}
\rho|_{z\rightarrow x_a}\sim (t_{3,a,0}+t_{3,a,1}(z-x^0_a))\bar{\partial}\chi_a(z,\bar{z}).
\end{equation}
The dual field $W$ has poles in the marked points
\beq{5.01}
W|_{z\rightarrow x_a}\sim
\frac{W_{-3,a}}{(z-x_a)^3}+\frac{W_{-2,a}}{(z-x_a)^2}+\frac{W_{-1,a}}{(z-x_a)}
+\ldots
\end{equation}
\begin{defi}
The space of ${\cal W}_3$-structure on $\Sigma_{g,n}$ is the subset of fields
in $\ti{\cal W}_3$ that satisfy the equations
\beq{5.1}
\kappa^2\partial^3\mu-(\bar{\partial}+\mu\partial+2\partial\mu)T=
\frac{2}{3}\kappa^2\partial^2(\partial^2-\f1{\kappa^2}T)\rho+\f1{\kappa}\partial(W\rho)-
\f1{\kappa}W(\mu-\partial\rho)
\end{equation}
\beq{5.2}
(\bar{\partial}+\rho\partial^2+(\mu+2\partial\rho)\partial+3\partial\mu+)W=
\kappa^3\partial[(\partial^2-\f1{\kappa^2}T)(\partial\mu-\frac{2}{3}(\partial^2-\f1{\kappa^2}T)\rho)].
\end{equation}
with fixed values of
\beq{5.2a}
{\bf T}_2=(T_{-2,1},\ldots,T_{-2,n})
\end{equation}
in (\ref{3.1b}) and
\beq{5.2b}
{\bf W}_2=(W_{-3,1},\ldots,W_{-3,n})
\end{equation}
in (\ref{5.01}).
\end{defi}
\begin{predl}
The equations (\ref{5.1}),(\ref{5.2}),
are the compatibility conditions for the linear system
\beq{5.11a}
(\kappa^3\partial^3+\kappa T\partial+W)\Psi=0,
\end{equation}
\beq{5.12a}
(\bar{\partial}+\frac{2}{3}(\partial^2-\f1{\kappa^2}T)\rho-\partial\mu+
(\mu-\partial\rho)\partial-\rho\partial^2)\Psi=0,
\end{equation}
where $\Psi$ is a section of $\Omega^{(-1,0)}(\Sigma_{g,n})$.
\end{predl}
The space ${\cal W}_3$ can be endowed with
the symplectic form
\beq{5.3}
\omega=\int_{\Sigma_{g,n}}(\delta T\delta\mu+\delta W\delta\rho).
\end{equation}
As for the projective connections the space ${\cal W}_3$ can be derived
from the space ${\cal N}_3$ of flat $\SL{3}$ connections
$$
{\cal N}_3=\{a,\bar{a}|F(a,\bar{a})=0\}
$$
over $\Sigma_{g,n}$ with the symplectic form (\ref{3.5}). In fact, (\ref{5.3})
is induced by (\ref{3.5}).
The flatness constraints generate the gauge symmetry of ${\cal N}_3$.
We assume that the matrix elements of the Lie algebra of the gauge group
are continuous in the marked points and, moreover,
\beq{5.3a}
x^{13}|_{z\rightarrow x_a}=O(z-x_a)^2,~~x^{23}|_{z\rightarrow x_a}=O(z-x_a).
\end{equation}
(see (\ref{3.5a})). In fact, (\ref{5.3a}) defines a subalgebra in the Lie
algebra of continuous gauge transformations.
The partial gauge fixing with respect to the parabolic subgroup
\cite{F}
\beq{ps}
P=\mathr{*}{*}{0}{*}{*}{0}{*}{*}{*}
\end{equation}
allows to pick up a special form of connections
\beq{5.4}
a=\mathr{0}{1}{0}{0}{0}{1}{W}{T}{0}
\end{equation}
with the prescribed behavior of $W$ and $T$ in neighborhoods of the marked
points. Then the exact form of the matrix $\bar{a}$ as well as (\ref{5.1}),
(\ref{5.2}) are extracted from the flatness condition.
The defining properties (\ref{5.3a}) of the gauge algebra allows to fix
the coefficients of highest poles (\ref{5.2a}),(\ref{5.2b}).
They define the value of the moment $m=F(a,\bar{a})$.
We generalize this procedure below.
The flatness
is the compatibility conditions for the linear system
$$
(\kappa\partial-a)\left(
\begin{array}{c}
\psi \\
\kappa\partial\psi \\
\kappa^2\partial^2\psi
\end{array}
\right )=0,\\
~~(\bar{\partial}-\bar{a})\left(
\begin{array}{c}
\psi \\
\kappa\partial\psi \\
\kappa^2\partial^2\psi
\end{array}
\right )=0,\\
$$
where $\psi\in\Omega^{(-1,0)}(\Sigma_{g,n})$. For the special form of $a$ (\ref{5.4})
and $\bar{a}$ it leads to (\ref{5.11a}),(\ref{5.12a}) in Proposition 4.1.
\bigskip
\noindent
{\bf 2. ${\cal W}_3^N$ structures.}\\
As in {\bf 3.1} consider $\SL{N}$-bundle $V$ over $\Sigma_{g,n}$ with
a fixed complex structure. Let $(A,\bar{A})$ be connections in $V$.
We generalize the ${\cal W}_3$ structure
(\ref{5.1}),(\ref{5.2}) taking into account the gauge degrees of freedom.
Define the space
$$
\ti{\cal W}_3^N=\ti{\cal W}_2^N\cup(W,\rho),
$$
where $\rho$ and $W$ satisfy (\ref{5.00}),(\ref{5.01}).
Introduce the following matrices
\beq{5.5}
\ti{T}=T-3(A^2+\kappa\partial A),
\end{equation}
\beq{5.6}
\ti{W}=W+TA-A^3-\kappa(A\partial A+\partial A^2)-\kappa^2\partial^2A.
\end{equation}
\begin{defi}
${\cal W}_3^N$ structure on $\Sigma_{g,n}$ is the subset of fields in
$\ti{\cal W}_3^N$
satisfying the following identities:
\beq{5.7}
\bar{\partial} \ti{W}-\kappa\partial f_5+\ti{W}f_1+\ti{T}f_3-3Af_5-f_7\ti{W}=0,
\end{equation}
\beq{5.8}
\bar{\partial} \ti{T}-\kappa\partial f_6+\ti{W}f_2+\ti{T}f_4-3Af_6-f_7\ti{T}=0,
\end{equation}
\beq{5.9}
\bar{\partial} A-\kappa\partial\bar{A}+[\bar{A},A]=0.
\end{equation}
\beq{5.9a}
A_{-1,a}=p_a,~(a=1,\ldots,n),
\end{equation}
\beq{5.9b}
T_{-2,a}=\frac{3<(p_a^0)^2>}{N}
~(a=1,\ldots,n).
\end{equation}
\beq{5.9c}
W_{-3,a}=\frac{<(p_a^0)^3>-3\kappa<(p_a^0)^2>}{N},~(a=1,\ldots,n).
\end{equation}
\end{defi}
The matrix coefficients $f_j$ have the form
\beq{f1}
f_1=-\frac{2}{3}(\partial^2-\f1{\kappa^2}T)\rho+\partial\mu-\bar{A}+\frac{2}{\kappa}A\partial\rho
-\f1{\kappa}\mu A-\frac{2}{\kappa^2}A^2\rho,
\end{equation}
\beq{f2}
f_2=-\f1{\kappa}(\mu-\partial\rho)-\frac{3}{\kappa^2}A\rho,
\end{equation}
\beq{f3}
f_3=\kappa\partial f_1-\f1{\kappa^2}\ti{W}\rho=
\end{equation}
$$
-\frac{2}{3}\kappa\partial(\partial^2-\f1{\kappa^2}T)\rho+\kappa\partial^2\mu-\kappa\partial\bar{A}+2\partial(A\partial\rho)
-\f1{\kappa^2}\ti{W}\rho-\partial(\mu A+\frac{2}{\kappa}A^2\rho),
$$
\beq{f4}
f_4=\kappa\partial f_2+f_1-\f1{\kappa^2}\ti{T}\rho=
\end{equation}
$$
\f1{3}(\partial^2-\f1{\kappa^2}T)\rho-\frac{1}{\kappa}A\partial\rho-\bar{A}+
\frac{1}{\kappa^2}A^2\rho-\f1{\kappa}\mu A,
$$
\beq{f5}
f_5=-\frac{\mu}{\kappa}\ti{W}+\kappa\partial f_3=
\end{equation}
$$
-\frac{2}{3}\kappa^2\partial^2(\partial^2-\f1{\kappa^2}T)\rho+\kappa^2\partial^3\mu-\kappa^2\partial^2\bar{A}+
2\kappa\partial^2(A\partial\rho)-\f1{\kappa}\partial(\ti{W}\rho)-\frac{\mu}{\kappa}\ti{W}
-\kappa\partial^2(\mu A +\frac{2}{\kappa}A^2\rho),
$$
\beq{f6}
f_6=-\frac{\mu}{\kappa}\ti{T}+\kappa\partial f_4+f_3=
\end{equation}
$$
-\frac{2}{3}\kappa\partial(\partial^2-\f1{\kappa^2}T)\rho+\kappa\partial^2\mu-\kappa\partial\bar{A}+
\partial(A\partial\rho)-\f1{\kappa^2}\ti{W}\rho-\frac{\mu}{\kappa}\ti{T}
+\frac{3}{\kappa}\partial(A^2)\rho
-2\partial(\mu A +\frac{2}{\kappa}A^2\rho),
$$
\beq{f7}
f_7=\frac{3}{\kappa}\mu A-\partial\mu+f_4=
\end{equation}
$$
\frac{2}{\kappa}\mu A-\partial\mu+\f1{3}(\partial^2-\f1{\kappa^2}T)\rho-
\frac{1}{\kappa}A\partial\rho-\bar{A}+\frac{1}{\kappa^2}A^2\rho.
$$
Note that (\ref{5.7}),(\ref{5.8}) is reduced to the standard
${\cal W}_3$ structure if $A=\bar{A}=0$, while (\ref{5.9}) is the flatness
of the bundle $V$.
\begin{predl}
The equations (\ref{5.7}),(\ref{5.8}),
(\ref{5.9}) are the compatibility conditions for the linear system
\beq{5.11}
(\kappa^3\partial^3+3A\kappa^2\partial^2+\kappa\ti{T}\partial+\ti{W})\Psi=0,
\end{equation}
\beq{5.12}
(\bar{\partial}-f_1-f_2\kappa\partial-\rho\partial^2)\Psi=0,
\end{equation}
where and $\Psi$ is a section of $\Omega^{(-1,0)}(\Sigma_{g,n},{\rm Aut}V)$.
\end{predl}
Again ${\cal W}_3^N$ has a natural symplectic form
\beq{5.10}
\omega=\int_{\Sigma_g}[-\frac{N}{\kappa}\delta T\delta\mu-\frac{N}{\kappa^2}\delta W\delta\rho+
3<\delta A,\delta\bar{A}>]+6\pi i\sum_{a=1}^n\delta<p_a,g_a^{-1}\delta g_a>.
\end{equation}
The proof of Proposition and the derivation of $\omega$ (\ref{5.10})
is based on the same procedure that
we already used in the ${\cal W}_2^N$ case. We consider
$\kappa\partial+{\cal A},~~\bar{\partial}+\bar{\cal A}$ connections in the
$(S(\GL{3}\otimes\GL{N}))$ vector bundle. The first component of
${\cal A},\bar{\cal A}$ acts on the sections of
$3$-jets of $\Omega^{(-1,0)}(\Sigma_{g,n})$, while the second component acts
on the sections of the bundle $\Omega^{(-1,0)}(\Sigma_{g,n},{\rm Aut}V)$.
We define the space ${\cal K}_3$ of connections $({\cal A},\bar{\cal A})$
and the set of coadjoint orbits (\ref{4.00}):
$$
({\cal A},\bar{\cal A},{\bf 1}_3\otimes{\bf p}).
$$
We assume that ${\cal A}$
satisfies additional restrictions:\\
$\bullet$ the ${\cal A}$ component are gauge equivalent
to the special form
\beq{4.60}
{\cal A}\sim DS\otimes{\bf 1}_N+{\bf 1}_3\otimes(-A),
\end{equation}
where
$$
DS=\mathr{0}{1}{0}{0}{0}{1}{W}{T}{0}
$$
$\bullet$ near the marked points $W$ has the form (\ref{5.01});\\
$\bullet$ near the marked points $T$ has the form (\ref{3.1a});\\
$\bullet$ near the marked points $A$ has the form (\ref{4.00}).
Proposition 4.1 follows from the following statement
\begin{predl}
${\cal W}_3^N$ is the subset of ${\cal K}_3$, satisfying the flatness condition
\beq{5.13}
\bar{\partial}{\cal A}-\kappa\partial\bar{\cal A}+[{\cal A},\bar{\cal A}]=0.
\end{equation}
\end{predl}
{\sl Proof.}\\
The symplectic form $\omega'$ on ${\cal K}_3$ has the similar form
as in the ${\cal K}_2$ case:
\beq{sf3}
\omega'=\int_{\Sigma_{g,n}}<<\delta{\cal A},\delta\bar{\cal A}>>+
6\pi i\sum_{a=1}^n\delta <p_a,g_a^{-1}\delta g_a>
\end{equation}
The gauge group symmetry of $\omega'$ is
$$
{\cal G}=\{{\rm Map}(\Sigma_{g,n}\rightarrow \SL{3N^2})\}.
$$
The action of ${\cal G}$ is the same as in (\ref{gt}).
For the matrix elements of Lie$({\cal G})$ we assume that they
are continuous in the marked points with additional restrictions
for the matrix elements $x^{13}_{\alpha\beta}$,$x^{23}_{\alpha\beta}$
$$
x^{13}_{\alpha\beta}\sim\delta_{\alpha\beta}O(z-x_a)^2+(1-\delta_{\alpha\beta})O(z-x_a)^3,
$$
$$
x^{23}_{\alpha\beta}\sim\delta_{\alpha\beta}O(z-x_a)+(1-\delta_{\alpha\beta})O(z-x_a)^2.
$$
Then for the dual space $y\in$Lie$^*({\cal G})$ we have
\beq{5.13b}
y^{31}_{\alpha\al}\sim\sum_{k\geq 2}b_k\partial^k \delta(x_a),~~
y^{31}_{\alpha\beta}\sim\sum_{k\geq 3}b_{k,\alpha,\beta}\partial^k \delta(x_a),~(\alpha\neq\beta).
\end{equation}
\beq{5.13a}
y^{32}_{\alpha\al}\sim\sum_{k\geq 1}b_k\partial^k \delta(x_a),~~
y^{32}_{\alpha\beta}\sim\sum_{k\geq 2}b_{k,\alpha,\beta}\partial^k \delta(x_a),~(\alpha\neq\beta).
\end{equation}
\beq{5.13c}
y^{JK}_{\alpha\beta}\sim\sum_{k\geq 0}b_{k,\alpha,\beta}\partial^k \delta(x_a),
~~(J\neq 3,~K\neq 1,2).
\end{equation}
We choose ${\cal A}$ in the form
\beq{5.17}
{\cal A}=\mathr{0}{E}{0}{0}{0}{E}{\ti{W}}{\ti{T}}{-3A}
\end{equation}
Substituting this form of ${\cal A}$ in (\ref{5.13}) we find $\bar{\cal A}$.
Solutions $\bar{\cal A}$ are
parametrized by the fields $\mu,\rho$ and the matrix $\bar{A}$.
Then
$$
\bar{\cal A}=
\mathr{f_1}{f_2}{-\rho/\kappa^2}{f_3}{f_4}{-\mu/\kappa}{f_5}{f_6}{f_7}.
$$
We set $f_1+f_4+f_7=-3\bar{A}$. This condition along with algebraic equation for
the blocks $(J,K),~J=1,2,~K=1,2,3$ of (\ref{5.13}) allows to find
$f_1,\ldots,f_7$ (\ref{f1})-(\ref{f7}).
The differential identities
arising in the last row $J=3,K=1,2,3$ lead to
(\ref{5.7}),(\ref{5.8}),(\ref{5.9}).
The behavior of the most singular terms near the
marked points (\ref{5.9a}),(\ref{5.9b}),(\ref{5.9c})
follows from the special form of Lie$^*({\cal G})$ (\ref{5.13a}),
(\ref{5.13b}),(\ref{5.13c}).
Moreover, $\omega'$ (\ref{sf3}) gives $\omega$ (\ref{5.10}) on ${\cal W}_3^N$
space.
Due to the special form of ${\cal A}$ and $\bar{\cal A}$
the compatible system of differential equations
$$
(\kappa\partial-{\cal A})\left(
\begin{array}{c}
\Psi \\
\kappa\partial\Psi \\
\kappa^2\partial^2\Psi
\end{array}
\right )=0,
$$
$$
(\bar{\partial}-\bar{\cal A})\left(
\begin{array}{c}
\Psi \\
\kappa\partial\Psi \\
\kappa^2\partial^2\Psi
\end{array}
\right )=0.
$$
is equivalent to (\ref{5.11}),(\ref{5.12}). $\Box$
\bigskip
\noindent
{\bf 3.${\cal W}^N_3$ and isomonodromic deformations.}\\
To get a nontrivial phase flow we should get rid of from the restrictive
constraints (\ref{5.7}),(\ref{5.8}),(\ref{5.9b}) and (\ref{5.9c}).
The remaining constraints generate the gauge subgroup
$$
{\cal G}^P=\left\{{\rm Map}(\Sigma_{g,n}\rightarrow P)\right\},
$$
where $P$ is the parabolic subgroup (\ref{ps}).
Thus, we come to the space ${\cal K}^N_3\subset{\cal K}_3$ of the
fields\\
$(W,\rho,T,\mu,A,\bar{A},p_a)$ with constraints (\ref{5.9}),(\ref{5.9a}).
The form $\omega$ (\ref{5.10}) is invariant under the following
transformations
\beq{5.18}
A\rightarrow A,~\bar{A}\rightarrow\bar{A}+\f1{\kappa}\xi_2A+\f1{\kappa^2}\xi_3A^2,~
T\rightarrow T,~\mu\rightarrow\mu+\xi_2,~\rho\rightarrow\rho+\xi_3,
\end{equation}
where $\xi_j\in\Omega^{(1-j,1)}(\Sigma_{g,n})$ with the same behavior near
the marked points as $\mu$ and $\rho$.
The moment constraints generated by these transformations are
$$
T=\frac{3}{2N}<A^2>,~~
W=\f1{N}<A^3>.
$$
Substituting $T$ and $W$ in $\omega$ (\ref{5.10}) we obtain
\beq{5.21}
\f1{3}\omega=\int_{\Sigma_g}<\delta A,\delta\bar{A}>+2\pi i\sum_{a=1}^n\delta<p_a,g_a^{-1}\delta g_a>
-\frac{1}{\kappa}\int_{\Sigma_{g,n}}<\delta A,A>\delta\mu-
\frac{1}{\kappa^2}\int_{\Sigma_{g,n}}<\delta A,A^2>\delta\rho.
\end{equation}
Since we don't fix the gauge of the transformations (\ref{5.18}),
the form $\omega$ becomes degenerated.
If one replace
$$
\bar{A}\rightarrow\bar{A}'=\bar{A}-\f1{\kappa}\mu A-\f1{2\kappa^2}\rho A^2
$$
$\omega$ takes the canonical form
$$
\f1{3}\omega=\int_{\Sigma_{g,n}}<\delta A,\delta\bar{A}'>
+2\pi i\sum_{a=1}^n\delta<p_a,g_a^{-1}\delta g_a>.
$$
It is invariant under the gauge transformations
\beq{5.22}
A\rightarrow fAf^{-1}+f\kappa\partial f^{-1},~~\bar{A}'\rightarrow f\bar{A}'f^{-1}+f\bar{\partial} f^{-1}.
\end{equation}
The moment equation resulting from this symmetry is
$$
F(A,\bar{A}'):=\bar{\partial} A-\kappa\partial\bar{A}+[\bar{A},A]=0,
$$
or in the original variables
\beq{5.24}
(\bar{\partial}+\partial\mu+\f1{2\kappa}\partial\rho A)A-\kappa\partial\bar{A}+[\bar{A},A]=0.
\end{equation}
In this case the constraints become nonlinear.
Now instead of infinite-dimensional space of smooth differentials
$\rho$ consider the finite dimensional space $H^1(\Sigma_{g,n},\Gamma^2)$.
It has dimension
\beq{5.30a}
l_3=\dim H^1(\Sigma_{g,n},\Gamma^2)=5g-5+2n.
\end{equation}
Expanding $\rho$ in the basis of $H^1(\Sigma_{g,n},\Gamma^2)$ we obtain
$
\rho=\sum_{s=1}^{l_3}t_{3,s}\rho_s^0.
$
Then $\omega$ (\ref{5.21}) gives
\beq{5.25}
\f1{3}\omega=\omega^0-\frac{1}{\kappa}\sum_{s=1}^{l_2}\delta H_{2,s}\delta t_{2,s}
-\frac{1}{2\kappa^2}\sum_{s=1}^{l_3}\delta H_{3,s}\delta t_{3,s},
\end{equation}
$$
H_{2,s}=\f1{2}\int_{\Sigma_{g,n}}<A^2>\mu_s^{(0)},~
H_{3,s}=\f1{3}\int_{\Sigma_{g,n}}<A^3>\rho_s^{(0)},
$$
where
$$
~~\omega^0=\int_{\Sigma_{g,n}}<\delta A,\delta\bar{A}>
+2\pi i\sum_{a=1}^n\delta<p_a,g_a^{-1}\delta g_a>.
$$
The equations of motion defining by $\omega$ are
\beq{5.28}
\partial_rA=0,~~~(r=(k,s),~k=2,3,~~\partial_r=\frac{\partial}{\partial t_{k,s}}),
\end{equation}
\beq{5.29}
\partial_r\bar{A}'=0.
\end{equation}
The solutions describe the free motion
\beq{5.30}
\bar{A}=\bar{A}_0+\f1{\kappa}A_0\sum_st_{2,s}+\frac{1}{2\kappa^2}A_0^2\sum_s t_{3,s}.
\end{equation}
Making use of the gauge symmetry we represent the fields as
$$
L=fAf^{-1}+f\kappa\partial f^{-1},
$$
$$
\bar{L}'=f\bar{A}'f^{-1}+f\bar{\partial} f^{-1},
$$
where
$$
M_r=f^{-1}\partial_rf.
$$
Thus we come to the finite-dimensional bundle
$$
{\cal P}_3^{red,N}=\{L,\bar{L},{\bf p},\mu,\rho\}.
$$
The bundle ${\cal P}_3^N$ has the same fibers ${\cal R}^{red}$ as the
bundle ${\cal P}_2^{red,N}$.
The equation (\ref{5.24}) takes the form
\beq{5.24a}
(\bar{\partial}+\partial\mu+\f1{2\kappa}\partial\rho L)L-\kappa\partial\bar{L}+[\bar{L},L]=0,
~~L|_{z\to x_a}=\frac{p_a}{z-x_a},
\end{equation}
where
$$
\bar{L}=\bar{L}'+\f1{\kappa}\mu L+\f1{2\kappa^2}\rho L^2.
$$
The equation of motion (\ref{5.28}),
on the reduced phase space $(L,\bar{L}')$ takes the form of the Lax
equation
\beq{5.34}
\partial_rL-\kappa\partial M_r+[M_r,L]=0.
\end{equation}
The second equation (\ref{5.29}) allows to define the matrices $M_r$ in the Lax
equation
\beq{5.35}
\partial_r\bar{L}'-\bar{\partial} M_r+[M_r,\bar{L}']=0.
\end{equation}
The equations (\ref{5.34}), (\ref{5.35}) with the flatness condition (\ref{5.24a})
are the compatibility conditions of the linear system
\beq{5.31}
(\kappa\partial+L)\psi=0,
\end{equation}
\beq{5.32}
(\bar{\partial}+\bar{L}')\psi=0,
\end{equation}
\beq{5.33}
(\kappa\partial_r+M_r)\psi=0.
\end{equation}
The last equation means that the monodromy matrix ${\cal M}$ of the linear
system (\ref{5.31}), (\ref{5.32}) on $\Sigma_{g,n}$ is independent with respect
to the ${\cal W}_3$ moduli $\mu_r,\rho_r$.
On the critical level $\kappa=0$ the systems pass into the Hitchin systems
with quadratic and cubic commuting integrals.
Consider a rational curve $\Sigma_{0,n}$ with $n$ fixed marked points
$x^0_1,\ldots,x^0_n$. According to (\ref{dim2}) and (\ref{5.30a}) we have
$8n-8$ times. Since in this case $\bar L=0$, (\ref{5.24a}) takes the form
$$
[\bar{\partial}+\partial\sum_{a=1}^nt_{2,a}\bar{\partial} \chi_a(z,\bar{z})
+\f1{2\kappa}\partial \sum_{a=1}^n (t_{3,a,0}+t_{3,a,1}(z-x^0_a))\bar{\partial}\chi_a(z,\bar{z}) L]L
=0.
$$
We remind that the times $t_{2,a}$ are related to the positions of the
marked points $(t_{2,a}=x_a-x_a^0)$.
If the $t_3$ times vanish then solution of this equation
$$
L=\sum_{a=1}^n\frac{p_a}{w-x_a},~~(w=z-\sum_{a=1}^nt_{2,a}\chi_a(z,\bar{z}))
$$
is the $L$ operator for the Schlesinger system.
For generic points $(t_{3,a}\neq 0)$ the
equation is nonlinear and its solutions are unknown.
\section{Conclusion}
\setcounter{equation}{0}
The generalization of the previous analysis on the arbitrary ${\cal W}_k^N$
structures is straightforward, though the explicit calculations on
intermediate steps become rather
long. In this section we present some general formulas.
Let $W_j,j=2,\ldots,k$ be the set of $j$-differentials on $\Sigma_{g,n}$ and
$\rho_j,j=2,\ldots,k$ are the dual objects, $(W_2=T,\rho_2=\mu)$. In addition
we have the gauge data $(A,\bar{A})$ and the coadjoint orbits in the
marked points $p_a,(a=1,\ldots,n)$
with the following behaviour
$$
W_j|_{z\rightarrow x_a}\sim
\frac{W_{-j,a}}{(z-x_a)^j}+\frac{W_{-j+1,a}}{(z-x_a)^{j-1}}+
+\ldots,
$$
$$
\rho_j|_{z\rightarrow x_a}\sim (t_{j,a,0}+t_{j,a,1}(z-x^0_a)+
\ldots+t_{j,a,j-2}(z-x_a)^{j-2})\bar{\partial}\chi_a(z,\bar{z}),
$$
and $A$ has poles of the first order (\ref{4.0}) with residues (\ref{4.6a}).
The highest order coefficients ${W_{-j,a}}$ are the linear combinations
of corresponding Casimirs up to the order $j$
$$
{W_{-j,a}}\sim\f1{N}(<(p_a^0)^j>+\ldots).
$$
Let $\Psi$ be a section of
$\Omega^{(-\frac{k-1}{2},0)}(\Sigma_{g,n},{\rm Aut}V)$. Then
the ${\cal W}_k^N$ structure on $\Sigma_{g,n}$ is the set of fields
$$
(W_j,\rho_j),~j=2,\ldots,k, (A,\bar{A}),p_a,a=1,\ldots,n,
$$
providing the consistency of the following linear system
$$
(\kappa^k\partial^k+kA\kappa^{k-1}\partial^{k-1}+\ldots+\ti{W}_k)\Psi=0,
$$
$$
(\bar{\partial}+\alpha_k\partial^{k-1}+\ldots+\alpha_1)\Psi=0.
$$
Here
$$
\ti{W}_k=W_k+AW_{k-1}+A^2W_{k-2}+\ldots+A^{k-2}W_2-A^k,
$$
and other coefficients can be read off from the flatness of the connections
${\cal A},\bar{\cal A}$ in the $\SL{kN}$ bundle with
$$
{\cal A}=DS\otimes {\bf I}_N+{\bf I}_k\otimes (-A)
$$
$$
DS=\left (
\begin {array}{cccccc}
0 & 1 & 0 & \cdots & & 0 \\
0 & 0 & 1 & \cdots & & 0 \\
\cdot & \cdot & \cdot & \cdots & &\cdot \\
0 & \cdot & \cdot & \cdots & 0 & 1\\
W_k & W_{k-1}& \cdot & \cdots & W_2 & 0
\end{array}
\right ).
$$
The symplectic form on ${\cal W}_k^N$
has the universal structure
\beq{7.1}
\omega=\int_{\Sigma_{g,n}}[k<\delta A,\delta\bar{A}>
-N\sum_{j=2}^k\f1{\kappa^{j-1}}\delta W_j\delta \rho_j]
+2k\pi i\sum_{a=1}^n\delta<p_a,g_a^{-1}\delta g_a>.
\end{equation}
We drop out details of calculations. Essentially they are the same as in
the case ${\cal W}_3^N$.
Again, the constraints on the fields
$W_j,\rho_j,j=2,\ldots,k$ can be discarded by the restriction to the
parabolic subgroup of the gauge transformations.
The form acquire the following Abelian symmetry
$$
\rho_j\rightarrow\rho_j+\xi_j,~(j=2,\ldots,k),~~
\bar{A}\rightarrow \bar{A}+\sum_{j=2}^k\frac{A^{j-1}\xi_j}{\kappa^{j-1}},
$$
where $\xi_j\in\Omega^{(-j,1)}(\Sigma_{g,n})$.
This symmetry is generated by the constraints
$$
W_j=\frac{k<A^j>}{Nj}.
$$
Substituting this expression in (\ref{7.1}) we obtain the symplectic form
\beq{7.2}
\f1{k}\omega=\int_{\Sigma_{g,n}}<\delta A,\delta\bar{A}>+
2\pi i\sum_{a=1}^n\delta<p_a,g_a^{-1}\delta g_a>
-\sum_{j=2}^k\f1{\kappa^{j-1}}\int_{\Sigma_{g,n}}<A^{j-1}\delta A> \delta \rho_j.
\end{equation}
Now we restrict the fields $\rho_j$ to the spaces
$H^1(\Sigma_{g,n},\Gamma^{j-1})$.
Let
$$
\rho_j=\sum_{s=1}^{l_j}t_{j,s}\rho_{j,s}^0, ~~~~
(l_j=\dim H^1(\Sigma_{g,n},\Gamma^{j-1})=(2j-1)(g-1)+(j-1)n)
$$
be the expansion of $\rho_j$ in the basis of $H^1(\Sigma_{g,n},\Gamma^{j-1})$.
Then we come to the following form \beq{7.3}
\f1{k}\omega=\omega^0-\sum_{j=2}^k\f1{\kappa^{j-1}}\sum_{s=1}^{l_j}\delta H_{j,s}\delta t_{j,s},
\end{equation}
where
$$
\omega^0=\int_{\Sigma_{g,n}}<\delta A,\delta\bar{A}>+2\pi i\sum_{a=1}^n\delta<p_a,g_a^{-1}\delta g_a>,
~~ H_{j,s}=\f1{j}\int_{\Sigma_{g,n}}<A^j>\rho_{j,s}^0.
$$
In terms of the field
$$
\bar{A}'=\bar{A}-\sum_{j=2}^k\f1{\kappa^{j-1}}A^{j-1} \rho_j
$$
it takes the canonical form
$$
\f1{k}\omega=\int_{\Sigma_g}[<\delta A,\delta\bar{A}'>+2\pi i\sum_{a=1}^n\delta<p_a,g_a^{-1}\delta g_a>.
$$
It is a free system with solutions
$$
A=A_0=const, ~~
\bar{A}({\bf t}=\bar{A}_0+
\sum_{j=2}^k\frac{A^{j-1}}{\kappa^{j-1}}
\sum_{s=1}^{l_j}t_{j,s}\rho_{j,s}^0,
$$
After the Hamiltonians reduction with respect to the gauge symmetry (\ref{5.22})
we come to the extended phase space
$$
{\cal P}_k^N=\{L,\bar{L},{\bf p},\rho_j,j=2,\ldots,k\}.
$$
The equations of motion are the same as in the ${\cal P}_3^N$ case
(\ref{5.34}),(\ref{5.35}), where
$$
\bar{L}'=\bar{L}\sum_{j=2}^{k}\f1{\kappa^{j-1}}\rho_{j}L^{j-1}.
$$
On the critical level we obtain the Hitchin systems with integrals
of order $j=2,\ldots,k$. In the case $k<N$ the number of integrals
is less then the dimension of the configuration space. For $k\geq N$ the
systems are completely integrable though the
integrals of order $j>N$ are not independent.
Thus, the distinguish case is $k=N$.
It is interesting to consider the limit $k\rightarrow\infty$.
Because the differential operators of an arbitrary order generate a
Lie algebra,
${\cal W}$ structure acquire the group-theoretical background as in
the ${\cal W}_2$ case. As we argued above the simultaneous limit
$k\rightarrow\infty,~N\rightarrow\infty$ is distinguish and can lead to the
interesting field theories.
{\bf Acknowledgments}\\
{\sl The work of A.L. is supported in part by grants RFFI-98-01-00344
and 96-15-96455 for support of scientific schools.
The work of M.O. is supported in part by grants RFFI-96-02-18046,
INTAS 96-518 and 96-15-96455 for support of scientific schools.
We are grateful to the Max-Planck-Institut f\"{u}r Mathemamatik in Bonn
for the hospitality, where this paper was prepared.
}
\small{
|
\section{Introduction} \label{intro}
The role of topological defects in symmetry-breaking phase transitions
cannot be underestimated. The prime example of a defect-driven
(equilibrium) transition is the superfluid-to-normal phase transition in
a $^4$He film, which can be understood as the unbinding of
vortex-antivortex pairs \cite{Berezinskii,KT73}. Since the very
existence of an ordered state and the presence of topological defects
are both manifestations of spontaneously broken symmetries, it doesn't
come as a surprise that various phase transitions can be understand in
terms of defects. To appreciate this point of view, note that defects
are regions, usually macroscopic in size, where the symmetry is broken
differently from that in the rest of the system. Often, the symmetry is
completely restored in the defect cores, meaning that they are in the
normal state. (In systems like superfluid $^3$He with a large symmetry
group, the defect cores may be ordered states themselves, but of a
different character.) Now, when defects with a normal-state core
proliferate, the ordered state gets disordered and converted to the
normal state.
In recent years also the dynamics of phase transitions and the
accompanying formation of defects have received much attention (see
Ref.\ \cite{nato} for reviews). An important role is played here by the
so-called time-dependent Ginzburg-Landau theory, which provides a
phenomenological approach. Ideally, one would like to start with a
microscopic model of the system under consideration and derive the
time-dependent Ginzburg-Landau theory by integrating out irrelevant
degrees of freedom, thus turning the phenomenological theory into an
effective one. For a superconductor this program, at least in the
time-independent case, has been carried out by Gorkov \cite{Gorkov}.
Starting from the microscopic theory of Bardeen, Cooper, and Schrieffer
(BCS) \cite{BCS}, he derived by means of a Green function method the
Ginzburg-Landau theory of superconductivity which had been proposed as a
phenomenological theory seven years before the BCS theory was
formulated \cite{GL}.
In the first part of this review paper, we reformulate Gorkov's
derivation (extended to include time dependence) in terms of the
functional-integral approach to quantum field theory. We will not only
study the weak-coupling BCS limit of loosely bound Cooper pairs, but
consider arbitrary values of the coupling constant $\lambda_0$,
including the limit $\lambda_0 \rightarrow -\infty$, where the fermions
form tightly bound pairs. When going from the weak-coupling to the
large-$\lambda_0$ limit, the dynamics changes from dissipative to
non-dissipative behavior.
In the second part of this review paper, we discuss the dual approach to the
time-independent Ginzburg-Landau theory. It is a formulation directly in
terms of magnetic vortices, in which the superconductor-to-normal is
described as a proliferation of vortices.
\subsection{Notation}
We adopt Feynman's notation and denote a spacetime point by $x=x_\mu
=(t,{\bf x})$, $\mu = 0,1, \cdots,d$, with $d$ the number of space
dimensions, while the energy $k_0$ and momentum ${\bf k}$ will be denoted by
$k=k_\mu = (k_0,{\bf k})$. The time derivative $\partial_0 =
\partial/\partial t$ and the gradient $\nabla$ are sometimes combined in a
single vector $\tilde{\partial}_\mu = (\partial_0, -\nabla)$. The tilde on
$\partial_\mu$ is to alert the reader for the minus sign appearing in the
spatial components of this vector. We define the scalar product $k \cdot x
= k_\mu x_\mu = k_\mu g_{\mu \nu} k_\nu = k_0 t - {\bf k} \cdot {\bf x}$,
with $g_{\mu \nu} = {\rm diag}(1,-1, \cdots,-1)$ and use Einstein's
summation convention. Because of the minus sign in the definition of the
vector $\tilde{\partial}_\mu$, it follows that $\tilde{\partial}_\mu a_\mu =
\partial_0 a_0 + \nabla \cdot {\bf a}$, with $a_\mu$ an arbitrary vector.
Integrals over spacetime are denoted by
$$
\int_{x} = \int_{t,{\bf x}} = \int \mbox{d} t \, \mbox{d}^d x,
$$
while those over energy and momentum by
$$
\int_k = \int_{k_0,{\bf k}} = \int \frac{\mbox{d} k_0}{2 \pi}
\frac{\mbox{d}^d k}{(2 \pi)^d}.
$$
When no integration limits are indicated, the integrals are assumed to
be over all possible values of the integration variables.
Natural units $\hbar = k_{\rm B} = 1$ are adopted throughout.
\section{Time-Dependent Ginzburg-Landau Theory} \label{sec:TDGL}
In this section, we derive the time-dependent Ginzburg-Landau theory for a
superconductor, starting from the microscopic BCS model. Following Ref.\
\cite{thesis}, we integrate out the fermionic degrees of freedom in favor of
a bosonic field describing the fermion pairs and use a derivative expansion
method. We wish to obtain the Ginzburg-Landau theory not only in the
weak-coupling BCS limit of loosely bound and overlapping Cooper pairs, but
for arbitrary values of the coupling constant, in particular in the limit
$\lambda_0 \rightarrow - \infty$ where the fermions form tightly bound pairs
\cite{Eagles,Leggett}. This crossover between the two limits has recently
been studied in detail \cite{DrZw,Haussmann,MRE,MPS}. Since in two space
dimensions this region becomes accessible by analytical methods, we shall,
unless stated otherwise, restrict ourselves to $d=2$ when carrying out
momentum integrals.
\subsection{BCS Model \label{sec:bcs}}
The BCS Lagrangian reads \cite{Popov}
\begin{eqnarray} \label{bcs:BCS}
{\cal L} &=& \psi^{\ast}_{\uparrow} [i\partial_0 - \xi(-i \nabla)]
\psi_{\uparrow}
+ \psi_{\downarrow}^{\ast} [i \partial_0 - \xi(-i \nabla)]\psi_{\downarrow}
- \lambda_0 \psi_{\uparrow}^{\ast}\,\psi_{\downarrow}
^{\ast}\,\psi_{\downarrow}\,\psi_{\uparrow} \nonumber \\
&=:& {\cal L}_{0} + {\cal L}_{\rm i},
\end{eqnarray}
where ${\cal L}_{\rm i} = - \lambda_0 \psi_{\uparrow}^{\ast} \,
\psi_{\downarrow}^{\ast}\,\psi_{\downarrow}\,\psi_{\uparrow}$ is a
contact interaction term, representing the effective attraction between
electrons with bare coupling constant $\lambda_0 < 0$, and ${\cal
L}_{0}$ is the remainder. In Eq.\ (\ref{bcs:BCS}), the field $\psi_{\uparrow
(\downarrow )}$ is an anticommuting field describing the electrons with
mass $m$ and spin up (down); $\xi(-i \nabla) = \epsilon(-i \nabla) -
\mu_0$, with $\epsilon(-i \nabla) = - \nabla^2/2m$, is the kinetic
energy operator with the bare chemical potential $\mu_0$ subtracted.
For computational convenience we introduce Nambu's notation and rewrite the
Lagrangian (\ref{bcs:BCS}) in terms of a two-component field
\begin{equation} \label{32}
\psi = \left( \begin{array}{c} \psi_{\uparrow} \\
\psi_{\downarrow}^{\ast} \end{array} \right) \:\:\:\:\:\:
\psi^{\dagger} = (\psi_{\uparrow}^{\ast},\psi_{\downarrow}).
\end{equation}
In order to integrate out the fermionic degrees of freedom, the
zero-temperature partition function represented as a functional integral
\begin{equation} \label{bcs:34}
Z = \int {\rm D} \psi^{\dagger} {\rm D} \psi \exp \left( i \int_x
\,{\cal L} \right),
\end{equation}
must be written in a form bilinear in the electron fields. This is achieved
by rewriting the quartic interaction term as a functional integral over
auxiliary fields $\Delta$ and $\Delta^*$:
\begin{eqnarray} \label{bcs:35}
\lefteqn{
\exp \left( -i \lambda_0 \int_x \psi_{\uparrow}^{\ast}
\, \psi_{\downarrow}^{\ast} \, \psi_{\downarrow}\, \psi_{\uparrow}
\right) = } \\
& & \int {\rm D} \Delta^* {\rm D} \Delta \exp \left[ -i
\int_x \left( \Delta^* \, \psi_{\downarrow}\,\psi_{\uparrow} +
\psi_{\uparrow}^{\ast} \, \psi_{\downarrow}^{\ast} \, \Delta -
\frac{1}{\lambda_0 } |\Delta|^2 \right) \right]. \nonumber
\end{eqnarray}
The field equation for $\Delta^*$,
\begin{equation} \label{bcs:del}
\Delta = \lambda_0 \psi_{\downarrow} \psi_{\uparrow},
\end{equation}
shows that the auxiliary field describes electron pairs. We will therefore
refer to it as pair field. The partition function thus becomes
\begin{eqnarray} \label{bcs:36}
\lefteqn{
Z = \int {\rm D} \psi^{\dagger} {\rm D} \psi \int {\rm D} \Delta^* {\rm
D} \Delta \; \exp\left(\frac{i}{\lambda_0} \int_x |\Delta|^2 \right)} \\
& & \times \exp \left[ i \int_x \, \psi^{\dagger} \left( \begin{array}{cc} i
\partial_{0} - \xi(-i \nabla) & -\Delta \\ -\Delta^* & i \partial_{0} +
\xi(-i \nabla)
\end{array} \right) \psi \right] \nonumber .
\end{eqnarray}
Changing the order of integration and performing the Gaussian integral over
the Grassmann fields, we obtain
\begin{equation} \label{bcs:37}
Z = \int {\rm D} \Delta^* {\rm D} \Delta \, \exp \left(i S_{\rm eff} [
\Delta^*, \Delta] + \frac{i}{\lambda_0}
\int_x |\Delta|^2 \right),
\end{equation}
with $S_{\rm eff}$ the one-loop effective action which, using the
identity ${\rm Det}(A) = \exp[{\rm Tr} \ln(A)]$, can be rewritten as
\begin{equation} \label{bcs:312}
S_{\rm eff}[\Delta^*, \Delta] = -i \, {\rm Tr} \ln \left(
\begin{array}{cc} p_{0} - \xi ({\bf p}) & -\Delta \\ -\Delta^* &
p_{0} + \xi ({\bf p})
\end{array}\right),
\end{equation}
where $p_0 = i \partial_0$ and $\xi({\bf p}) = \epsilon({\bf p}) - \mu_0$,
with $\epsilon({\bf p}) = {\bf p}^2/2m$.
The trace Tr appearing here needs some explanation. Explicitly, it is
defined as
\begin{equation} \label{bcs:explicit}
S_{\rm eff} = -i {\rm Tr} \, \ln \left[K(p,x) \right] = -i {\rm tr} \ln\left[
K(p,x) \delta (x - y)\bigr|_{y = x} \right],
\end{equation}
where the trace tr is the usual one over discrete indices. We
abbreviated the matrix appearing in (\ref{bcs:312}) by $K(p,x)$ so as to
cover the entire class of actions of the form
\begin{equation}
S = \int_x \psi^\dagger(x) K(p,x) \psi(x).
\end{equation}
The delta function in (\ref{bcs:explicit}) arises because $K(p,x)$ is
obtained as a second functional derivative of the action
\begin{equation}
\frac{\delta^{2} S}{\delta \psi^\dagger(x) \, \delta \psi(x)} =
K(p,x) \, \delta (x - y) \bigr|_{y = x},
\end{equation}
each of which gives a delta function. Since the action has only one
integral $\int_x$ over spacetime, one delta function remains. Because it is
diagonal, it may be taken out of the logarithm and the effective action
(\ref{bcs:explicit}) can be written as
\begin{eqnarray} \label{bcs:Trexplicit}
S_{\rm eff} &=& -i {\rm tr} \, \int_x
\ln \left[ K(p,x) \right]
\delta (x - y) \bigr|_{y = x} \nonumber \\ &=&
-i {\rm tr} \, \int_x \int_k {\rm e}^{i k \cdot x} \, \ln \left[ K(p,x)
\right] {\rm e}^{-i k \cdot x}.
\end{eqnarray}
In the last step, we used the integral representation of the
delta function:
\begin{equation}
\delta (x) = \int_k {\rm e}^{-i k \cdot x},
\end{equation}
shifted the exponential function $\exp (i k \cdot y)$ to the left, which
is justified because the derivative $p_\mu$ does not operate on it, and,
finally, set $y_\mu$ equal to $x_\mu$. We thus see that the trace Tr in
(\ref{bcs:explicit}) stands for the trace over discrete indices as well
as the integration over spacetime and over energy and momentum. The
integral $\int_k$ arises because the effective action
(\ref{bcs:explicit}) is a one-loop result with $k_\mu$ the loop energy
and momentum.
In the mean-field approximation, the functional integral (\ref{bcs:37}) is
approximated by the saddle point:
\begin{equation} \label{bcs:38}
Z = \exp \left(i S_{\rm eff}
[ \Delta^*_{\rm mf}, \Delta_{\rm mf} ] + \frac{i}{\lambda_0} \int_x
|\Delta_{\rm mf}|^2 \right),
\end{equation}
where $\Delta_{\rm mf}$ is the solution of the mean-field equation
\begin{equation} \label{bcs:gap}
\frac{\delta S_{\rm eff} }{\delta \Delta^*(x)
} = - \frac{1}{\lambda_0} \Delta (x) .
\end{equation}
If we assume the system to be spacetime independent, so that $\Delta_{\rm
mf}(x) = \bar{\Delta}$, Eq.\ (\ref{bcs:gap}) yields the celebrated BCS gap
equation \cite{BCS}:
\begin{eqnarray} \label{bcs:gape}
\frac{1}{\lambda_0} &=& - i \int_k \frac{1}{k_{0}^{2} - E^{2}(k) + i \eta}
\nonumber \\ &=& - \frac{1}{2} \int_{\bf k} \frac{1}{E({\bf k})},
\end{eqnarray}
where $\eta$ is an infinitesimal positive constant that is to be set to
zero at the end of the calculation, and
\begin{equation} \label{bcs:spec}
E({\bf k}) = \sqrt{\xi^2({\bf k}) + |\bar{\Delta}|^2}
\end{equation}
is the spectrum of the elementary fermionic excitations.
\subsection{BCS to BEC}
For a constant pair field, the effective action can be calculated in
closed form. Writing
\begin{equation}
\left(
\begin{array}{cc} k_{0} - \xi ({\bf k}) & -\bar{\Delta} \\ -\bar{\Delta}^* &
k_{0} + \xi ({\bf k}) \end{array}\right) = \left(
\begin{array}{cc} k_{0} - \xi ({\bf k}) & 0 \\ 0 &
k_{0} + \xi ({\bf k}) \end{array}\right) - \left(
\begin{array}{cc} 0 & \bar{\Delta} \\ \bar{\Delta}^* & 0 \end{array}\right),
\end{equation}
and expanding the second logarithm in a Taylor series, we recognize the
form
\begin{eqnarray}
\lefteqn{S_{\rm eff}[\bar{\Delta}^*, \bar{\Delta}] = } \nonumber \\ &&
-i \, {\rm Tr} \ln \left(
\begin{array}{cc} k_{0} - \xi ({\bf k}) & 0 \\ 0 &
k_{0} + \xi ({\bf k}) \end{array}\right) - i \, {\rm Tr} \ln
\left(1 - \frac{|\bar{\Delta}|^2}{k_0^2 - \xi^2({\bf k})} \right),
\end{eqnarray}
where we ignored an irrelevant constant. The integral over the loop
energy $k_0$ can be carried out to yield for the effective Lagrangian
\begin{equation} \label{bcs:closed}
{\cal L}_{\rm eff} = \int_{\bf k} \left[ E({\bf k}) - \xi({\bf k})
\right].
\end{equation}
To this one-loop result we have to add the tree term
$|\bar{\Delta}|^2/\lambda_0$. Expanding $E({\bf k})$ in a Taylor series, we
see that the effective Lagrangian also contains a term quadratic in
$\bar{\Delta}$. This term diverges in the ultraviolet. To render the
theory finite to this order, we have to introduce a renormalized coupling
constant $\lambda$ defined by:
\begin{equation} \label{bcs:reng}
\frac{1}{\lambda} = \frac{1}{\lambda_0} + \frac{1}{2} \int_{\bf k}
\frac{1}{|\xi({\bf k})|}.
\end{equation}
To this order in the loop expansion there is no renormalization of the
chemical potential, so that we can write $\mu = \mu_0$. We regularize
the diverging integral in Eq. (\ref{bcs:reng}) by introducing a momentum
cutoff $\Lambda$. In, for example, $d=3$, we then obtain
\begin{equation} \label{bcs:ren}
\frac{1}{\lambda} = \frac{1}{\lambda_0} + \frac{m}{2 \pi^2}
\Lambda,
\end{equation}
where we omitted the (irrelevant) finite part of the integral. It should be
remembered that the bare coupling constant $\lambda_0$ is negative, so that
the interaction between the fermions is attractive. We can distinguish two
limits. One, the famous weak-coupling BCS limit, which is obtained by
taking the bare coupling constant to zero, $\lambda_0 \rightarrow 0^-$.
Second, the limit which is obtained by letting $\lambda_0 \rightarrow -
\infty$. In this limit, the two-particle interaction is such that the
fermions form tightly bound pairs of mass $2m$ \cite{Eagles,Leggett}.
To explicate this so-called Bose-Einstein condensation (BEC) limit in
$d=2$, we swap the bare coupling constant for a more convenient
parameter, namely the binding energy $\epsilon_a$ of a fermion pair in
vacuum \cite{RDS}. Both parameters characterize the strength of the
contact interaction. To see the connection between the two, let us
consider the Schr\"odinger equation for the problem at hand. In reduced
coordinates, it reads
\begin{equation}
\left[- \frac{\nabla^2}{m} + \lambda_0 \, \delta({\bf x}) \right] \psi({\bf
x}) = - \epsilon_a \psi({\bf x}),
\end{equation}
where the reduced mass is $m/2$ and the delta-function potential, with
$\lambda_0 < 0$, represents the attractive contact interaction ${\cal
L}_{\rm i}$ in the BCS Lagrangian (\ref{bcs:BCS}). We stress that this is a
two-particle problem in vacuum; it is not the famous Cooper problem of two
interacting fermions on top of a filled Fermi sea. The equation is most
easily solved by Fourier transforming it. This yields the bound-state
equation
\begin{equation}
\psi({\bf k}) = - \frac{\lambda_0}{{\bf k}^2/m + \epsilon_a} \psi(0),
\end{equation}
or
\begin{equation}
- \frac{1}{\lambda_0} = \int_{\bf k} \frac{1}{{\bf k}^2/m + \epsilon_a} =
\frac{1}{2} \nu(0) \ln \! \left( \frac{2 \epsilon_\Lambda}{\epsilon_a}
\right),
\end{equation}
where $\nu(0) = m/2 \pi$ is the two-dimensional density of states (per
spin degree of freedom), and $\epsilon_\Lambda = \Lambda^2/2m$. This
equation allows us to replace the bare coupling constant $\lambda_0$ with
the binding energy $\epsilon_a$. When substituted in the gap equation
Eq.\ (\ref{bcs:gape}), the latter becomes
\begin{equation} \label{bcs:reggap}
\int_{\bf k} \frac{1}{{\bf k}^2/m + \epsilon_a} = \frac{1}{2}
\int_{\bf k} \frac{1}{E({\bf k})}.
\end{equation}
By inspection, it is easily seen that this equation has a solution
\cite{Leggett}
\begin{equation} \label{comp:self}
\bar{\Delta} \rightarrow 0, \;\;\;\;\; \mu_0 \rightarrow - \tfrac{1}{2}
\epsilon_a,
\end{equation}
with a negative chemical potential. This is the strong-coupling BEC
limit. To appreciate the physical significance of the specific value
found for the chemical potential in this limit, we note that the
spectrum $E_{\rm b}({\bf q})$ of the two-fermion bound state measured
relative to the pair chemical potential $2\mu_0$ reads
\begin{equation}
E_{\rm b}({\bf q}) = - \epsilon_a + \frac{{\bf q}^2}{4m} -2 \mu_0.
\end{equation}
The negative value for $\mu_0$ found in (\ref{comp:self}) is precisely the
condition for a Bose-Einstein condensation of the composite bosons in the
${\bf q} = 0$ state---whence the name BEC limit.
Since there are two unknowns contained in the theory, viz.\
$\bar{\Delta}$ and $\mu$, a second equation is needed to determine these
variables in the mean-field approximation \cite{Leggett}. It is
provided by the requirement that the average fermion number $N$, which
is obtained by differentiating the effective action (\ref{bcs:312}) with
respect to $\mu$
\begin{equation}
N = \frac{\partial S_{\rm eff}}{\partial \mu},
\end{equation}
be fixed. If the system is spacetime independent, this reduces to
\begin{equation} \label{bcs:n}
\bar{n} = - i\, {\rm tr} \int_k \, G(k) \tau_3,
\end{equation}
where $\bar{n}=N/V$, with $V$ the volume of the system, is the constant
fermion number density, $\tau_3$ is the diagonal Pauli matrix in Nambu space,
\begin{equation}
\tau_3 = \left(
\begin{array}{cr} 1 & 0 \\ 0 & -1
\end{array} \right),
\end{equation}
and $G(k)$ is the Green function,
\begin{eqnarray} \label{bcs:prop}
G(k) &=&
\left( \begin{array}{cc} k_0 - \xi ({\bf k})
& -\bar{\Delta} \\ -\bar{\Delta}^* & k_0 + \xi ({\bf k})
\end{array} \right)^{-1} \\ &=&
\frac{1}{k_0^2 - E^2({\bf k}) + i \eta }
\left( \begin{array}{cc} k_{0} \, {\rm e}^{i k_0 \eta } + \xi
({\bf k}) &
\bar{\Delta} \\ \bar{\Delta}^* & k_{0} \, {\rm e}^{-i k_0 \eta}- \xi
({\bf k}) \end{array} \right). \nonumber
\end{eqnarray}
Here, $\eta$ is an infinitesimal positive constant that is to be set to zero
at the end of the calculation. The exponential functions in the diagonal
elements of the Green function are an additional convergence factor needed in
nonrelativistic theories \cite{Mattuck}. If the integral over the loop
energy $k_0$ in the particle number equation (\ref{bcs:n}) is carried out,
it takes the familiar form
\begin{equation} \label{bcs:ne}
\bar{n} = \int_{\bf k} \left(1 - \frac{\xi({\bf k})}{E({\bf k})} \right).
\end{equation}
The two equations (\ref{bcs:gape}) and (\ref{bcs:n}) determine
$\bar{\Delta}$ and $\mu$.
\subsection{Derivative Expansion \label{bcs:s5}}
When the pair field $\Delta_{\rm}$ is spacetime-dependent, the
integrals in (\ref{bcs:312}) cannot be evaluated in closed form because the
logarithm contains energy-momentum operators and spacetime-dependent
functions in a mixed order. To disentangle the integrals we have to resort
to a derivative expansion \cite{FAF} in which the logarithm is expanded in a
Taylor series. Each term contains powers of the energy-momentum operator
$p_\mu$ which acts on every spacetime-dependent function to its right. All
these operators are shifted to the left by repeatedly applying the identity
\begin{equation}
f(x) p_\mu g(x) = (p_\mu - i \tilde{\partial}_\mu) f(x) g(x),
\end{equation}
where $f(x)$ and $g(x)$ are arbitrary functions of spacetime and the
derivative $\tilde{\partial}_\mu = (\partial_0,-\nabla)$ acts {\it only} on
the next object to the right. One then integrates by parts, so that all the
$p_\mu$'s act to the left where only a factor $\exp(i k \cdot x)$ stands.
Ignoring total derivatives and taking into account the minus signs that
arise when integrating by parts, one sees that all occurrences of $p_\mu$
(an operator) are replaced with $k_\mu$ (an integration variable). The
exponential function $\exp(-i k \cdot x)$ can at this stage be moved to the
left where it is annihilated by the function $\exp(i k \cdot x)$. The
energy-momentum integration can now in principle be carried out and the
effective action be cast in the form of an integral over a local density
${\cal L}_{\rm eff}$:
\begin{equation}
S_{\rm eff} = \int_x {\cal L}_{\rm eff}.
\end{equation}
This is in a nutshell how the derivative expansion works \cite{FAF}.
To apply it to the BCS model and derive the Ginzburg-Landau theory we use
the following decomposition in (\ref{bcs:312}):
\begin{equation} \label{313}
\left( \begin{array}{cc} p_{0} - \xi ({\bf p}) & -\Delta \\
-\Delta^* & p_{0} + \xi ({\bf p}) \end{array} \right) =
G_0^{-1} \left[ 1 - G_0 \,
\left( \begin{array}{cc} 0 & \Delta \\
\Delta^* & 0 \end{array} \right) \right],
\end{equation}
where $G_0$ is the Green function (\ref{bcs:prop}) with $\bar{\Delta}=0$.
Apart from an irrelevant constant, this leads to the expression for the
effective action
\begin{eqnarray} \label{316}
S_{\rm eff} &=& - i \, {\rm Tr} \, \ln \left[ 1 - G_0 \, \left(
\begin{array}{cc} 0 & \Delta \\ \Delta^* & 0 \end{array}\right)
\right] \nonumber \\ &=& i \, \mbox{Tr} \, \sum_{\ell=1}^{\infty} \,
\frac{1}{\ell}\, \left[ G_0 \, \left(\begin{array}{cc} 0 & \Delta \\
\Delta^* & 0
\end{array}\right) \right]^\ell =: \sum_{\ell=1}^{\infty}
S_{\rm eff}^{(\ell)}.
\end{eqnarray}
For $\ell=1$, the trace over the $2 \times 2$ matrix immediately yields zero.
In a similar fashion all terms $S_{\rm eff}^{(\ell)}$, with $\ell$ odd, give
zero. For the quadratic term, we obtain
\begin{equation} \label{317}
S_{\rm eff}^{(2)} = i \, {\rm Tr} \, \frac{1}{p_{0} + \xi ({\bf p})} \,
\Delta^* \, \frac{1}{p_{0} - \xi ({\bf p})} \, \Delta,
\end{equation}
where we recall the definition of the derivative $p_\mu$ as operating on
everything that appears to the right. Applying the derivative expansion
rules outlined above, we can cast the quadratic term in the effective
action in the form
\begin{equation} \label{319}
S_{\rm eff}^{(2)} = i \, {\rm Tr} \, \frac{1}{k_{0} + \xi ({\bf k})} \,
\frac{1}{k_{0} - i \, \partial_{0} - \xi ({\bf k} + i \nabla)} \, \Delta^*
\Delta.
\end{equation}
Usually, the quartic terms are included in the Ginzburg-Landau theory
without derivatives. We then can treat $p_0 \pm \xi ({\bf p})$ as a
c-number and
\begin{equation} \label{320}
S_{\rm eff}^{(4)} = \frac{i}{2} \, {\rm Tr} \, \frac{1}{[\, k_{0}^2 - \xi^2
({\bf k})\, ]^{2}} \, |\Delta|^4.
\end{equation}
We may truncate the series in (\ref{316}) here, provided the pair field
$\Delta$ is small as is the case in the vicinity of the phase
transition.
To include the temperature $T$ in the theory, we adopt the
imaginary-time approach to thermal field theory \cite{Rivers,Kapusta}.
Very briefly, it can be derived from the corresponding quantum field
theory at zero temperature simply by going over to imaginary times, $t
\rightarrow -i \tau$, and substituting
\begin{equation} \label{fun:sub}
\int \frac{{\rm d} k_{0}}{2\pi}\,g(k_{0})\rightarrow i\, \beta^{-1}
\sum_{n} \, g( i\,\omega_{n}),
\end{equation}
where $g$ is an arbitrary function, while $\omega_n$ denote the Matsubara
frequencies,
\begin{equation} \label{fun:matb}
\omega_n = \left\{ \begin{array}{ll} \pi \beta^{-1} 2n, &
\;\;\;\; \mbox{(bosonic)} \\ \pi \beta^{-1}(2n + 1), & \;\;\;\; \mbox{(fermionic)}
\end{array} \right.
\end{equation}
with $n$ an integer and $\beta= 1/T$. With these rules, the Minkowski
action $S$ goes over into
\begin{equation} \label{fun:MtE}
S = \int_{x} {\cal L} (t,{\bf x}) \rightarrow -i \int_{0}^{\beta} {\rm d} \tau
\int_{\bf x} {\cal L} (-i\tau, {\bf x}) =: i S^{\rm E},
\end{equation}
where the superscript E on the action at the right-hand side is to indicate
that it pertains to Euclidean rather than to Minkowski spacetime.
For the case at hand we obtain in this way the finite-temperature action:
\begin{equation} \label{51}
S_{\rm eff}^{\rm E} = {\rm Tr}\, \left[ \frac{1}{i \omega_{n} + \xi ({\bf
k})} \, \frac{1}{i\omega_{n} + \partial_{\tau} - \xi ({\bf k} + i \nabla)}
\Delta^* \Delta + \frac{1}{2} \frac{1}{[\omega_{n}^2 + \xi^2 ({\bf
k})]^{2}} |\Delta|^4 \right] ,
\end{equation}
where the trace at finite temperature reads explicitly
\begin{equation}
{\rm Tr} = \int_{0}^{\beta} {\rm d} \tau \int {\rm d}^dx \frac{{\rm d}^d
k}{(2\pi)^d} \, \beta^{-1} \sum_{n}.
\end{equation}
Let us for the moment consider only the time-independent part of the
effective action (\ref{51}) and expand in gradients. The sums over the
Matsubara frequencies are carried out with the help of the formulas
\begin{eqnarray} \label{sums}
\beta^{-1} \sum_n \frac{1}{\omega_n^2 + \xi^2} &=& \frac{1}{2}
\frac{X}{\xi} \nonumber \\ \beta^{-1} \sum_n \frac{1}{(\omega_n^2 +
\xi^2)^2} &=& -\frac{1}{4 \xi} \frac{\partial}{\partial \xi}
\left(\frac{X}{\xi}\right) \nonumber \\ \beta^{-1} \sum_n \frac{1}{i
\omega_n + \xi}\frac{1}{(i \omega_n - \xi)^2} &=& -\frac{1}{4}
\frac{\partial}{\partial \xi} \left(\frac{X}{\xi}\right) \nonumber \\
\beta^{-1} \sum_n \frac{1}{i \omega_n + \xi}\frac{1}{(i \omega_n - \xi)^3}
&=& - \frac{X}{8 \xi^3} -\frac{\beta}{16 \xi} \frac{\partial}{\partial \xi}
\left(\frac{Y}{\xi}\right) ,
\end{eqnarray}
where $X$ and $Y$ abbreviate the functions
\begin{equation}
X = \tanh(\beta \xi/2), \;\;\;\;\;\; Y = 1/\cosh^2(\beta \xi/2).
\end{equation}
As was first shown by Drechsler and Zwerger \cite{DrZw}, most of the
momentum integrals can be performed analytically for arbitrary values of the
coupling constant in two space dimensions. In this way, they arrived at the
time-independent Ginzburg-Landau theory describing the crossover from the
weak-coupling BCS limit to the strong-coupling BEC limit:
\begin{equation} \label{general}
{\cal L}_{\rm eff} = c \Delta^* \nabla^2 \Delta + a |\Delta|^2 -
\tfrac{1}{2} b |\Delta|^4,
\end{equation}
with the coefficients
\begin{eqnarray} \label{a}
a = \frac{\nu(0)}{2} \!\!\!\! & \Biggl[ & \!\!\!\! 2 \ln\left(\frac{4 {\rm
e}^\gamma}{\pi} \right) \theta(\mu) + \ln(\beta \epsilon_a/4) + \ln(\beta
|\mu|/2) \tanh(\beta \mu/2) \nonumber \\ & & \!\!\!\! + {\rm sgn}(\mu)
\int_{\beta |\mu|/2}^\infty {\rm d} x \, \ln(x) Y(x) \Biggr],
\end{eqnarray}
\begin{equation} \label{b}
b = \frac{\nu(0)}{4} \left[ \frac{7 \zeta(3)}{2 \pi^2} \beta^2 \theta(\mu) -
\frac{1}{\mu^2} \tanh(\beta \mu/2) + {\rm sgn}(\mu) \frac{\beta^2}{4}
\int_{\beta |\mu|/2}^\infty {\rm d} x \frac{X(x)}{x^3} \right],
\end{equation}
and
\begin{equation} \label{c}
c = \frac{\nu(0)}{8 m} \left[ \frac{7 \zeta(3)}{2 \pi^2} \beta^2 \mu
\theta(\mu) + \frac{\beta^2}{4} |\mu| \int_{\beta |\mu|/2}^\infty {\rm d} x
\frac{X(x)}{x^3} \right].
\end{equation}
Here, we introduced the integration variable $x = \beta \xi/2$, $\gamma =
0.577216 \cdots$ is Euler's constant, $\theta(x)$ is the Heaviside unit step
function, ${\rm sgn}(x)$ the sign function, and $\zeta(x)$ Riemann's zeta
function, with $\zeta(3)= 1.20206 \cdots$. (The second term at the
right-hand side of Eq. (\ref{b}) differs slightly from the corresponding
term in Ref.\ \cite{DrZw}, but is consistent with a later work by Zwerger
and collaborator \cite{SZ}.)
When studying the time-dependence of the effective action (\ref{51}),
special care has to be taken with analytic continuation. We follow S\'a de
Melo, Randeria, and Engelbrecht \cite{MRE} and first analytic continue,
using the formula
\begin{equation} \label{ana}
\beta^{-1} \sum_n \frac{1}{i \omega_n + \xi}\frac{1}{i \omega_n - i \omega_l
- \xi} = - \left[{\rm P} \frac{1}{2 \xi + q_0} - i \pi \delta(2 \xi +
q_0)\right] X,
\end{equation}
before expanding in time derivatives. (In Ref.\ \cite{DrZw}, the
expansion was done first, leading to results which are not consistent
with those known in the BCS limit \cite{Schmid,AT}.) In Eq.\
(\ref{ana}), ${\rm P}$ stands for the principal part, while $\omega_l$
is a bosonic Matsubara frequency, which at the right-hand side is
analytic continued to the real axes by replacing $i\omega_l$ with $q_0 +
i \eta$. In this way, we find for the dynamic part (in
Minkowski spacetime)
\begin{equation} \label{dynamic}
{\cal L}_{\rm dyn} = [Q'(i \partial_0) - i \pi Q''(i \partial_0)] \Delta^*
\Delta,
\end{equation}
with
\begin{eqnarray}
Q'(q_0) &=& - {\rm P} \int_{\bf k} \frac{q_0}{2 \xi(2 \xi + q_0)} X
\nonumber \\ Q''(q_0) &=& \tfrac{1}{2} \nu(0) \tanh(\beta q_0/4) \theta(\mu
- q_0/2).
\end{eqnarray}
The most important result to be noted here is that the low-energy dynamics
is dissipative when the chemical potential $\mu$ is positive \cite{MRE}.
This is because the pairs can break up and decay into a continuum of
fermionic excitations. On the other hand, for negative values of $\mu$,
where the pairs are more tightly bound, the time-dependent Ginzburg-Landau
theory describes a purely propagating pair mode. To investigate this point
further, let us consider the two limits in detail.
\subsection{BCS limit}
In the weak-coupling BCS limit, where the chemical potential is well
approximated by the Fermi energy $\mu = k_{\rm F}^2/2m$, with $k_{\rm F}$
the Fermi momentum, and $\epsilon_a/\epsilon_{\rm F} \rightarrow 0$, we
recover the standard result (adjusted for the reduced space dimensionality)
\cite{AT}
\begin{equation} \label{521}
{\cal L}_{\rm eff} = \nu(0) \left\{\Delta^* \left[ \ln \!
\left(\frac{T_0}{T}\right) - \frac{\pi}{8 T_0} \partial_{0} + \frac{3
\xi_{0}^{2} }{v_{\rm F}^{2}} \left(\partial_{0}^{2} + \frac{v_{\rm F}^2}{2}
\nabla^2\right) \right] \Delta - \frac{3 \xi_{0}^{2} }{v_{\rm F}^2}
|\Delta|^4 \right\}
\end{equation}
with $v_{\rm F}= k_{\rm F}/m$ the Fermi velocity, $\xi_0$ the BCS
correlation length
\begin{equation}
\xi_{0}^2 = \frac{7 \zeta(3) }{48 \pi^2} \frac{v_{\rm F}^2}{T_0^2},
\end{equation}
and $T_0$ the BCS transition temperature \cite{DrZw}
\begin{equation}
T_0 = \frac{{\rm e}^\gamma}{\pi} (2 \mu \epsilon_a)^{1/2}
\end{equation}
expressed in terms of the binding energy $\epsilon_a$. Comparing the
two terms in Eq.\ (\ref{521}) involving time derivatives, we recognize a
series expansion in powers of $q_0/T$; the terms without derivatives
constitute an expansion in powers of $|\Delta|/T$ \cite{bosepaper}. It
therefore follows that Eq.\ (\ref{521}) represents the first terms in a
high-temperature expansion. Since we neglected time derivatives in the
quartic and higher order terms, we are implicitly assuming that $q_0 >>
|\Delta|$ \cite{AT}. The time-dependent Ginzburg-Landau theory
(\ref{521}) obtained in the weak-coupling BCS limit is of the
dissipative type often used in the context of defect formation in a
symmetry-breaking phase transition (for reviews see Ref.\ \cite{nato}).
For applications in the context of superconductivity see, for example,
Refs.\ \cite{DG,Tinkham,Crisan}.
\subsection{BEC Limit}
In the strong-coupling BEC limit, the general form of the effective
action (\ref{general}) and (\ref{dynamic}) reduces to \cite{DrZw}
\begin{equation} \label{GP}
{\cal L}_{\rm eff} = \hat{\Delta}^* \left( i \partial_0 + \mu_{\rm b} +
\frac{\nabla^2}{4 m} \right) \hat{\Delta} - \lambda_{\rm b}
|\hat{\Delta}|^4,
\end{equation}
where we introduced a rescaled pair field
\begin{equation}
\hat{\Delta} = \left(\frac{\nu(0)}{4 |\mu|}\right)^{1/2} \Delta.
\end{equation}
The effective theory (\ref{GP}) is precisely of the form of a
Gross-Pitaevski theory \cite{GrPi}, describing a weakly interacting
composite Bose gas with a mass $2m$ (as expected), a small chemical
potential which vanishes on approaching the critical temperature from
below [see Eq.\ (\ref{comp:self})]
\begin{equation}
\mu_{\rm b} = 2 |\mu| \, \ln \left(\frac{\epsilon_a}{2|\mu|}\right),
\end{equation}
or using Eq.\ (\ref{bcs:reggap}),
\begin{equation}
\mu_{\rm b} = \frac{|\bar{\Delta}|^2}{\epsilon_a},
\end{equation}
and a repulsive contact interaction
\begin{equation}
\lambda_{\rm b} = \frac{1}{\nu(0)}
\end{equation}
independent of the binding energy $\epsilon_a$ characterizing the
interaction between the electrons. This is special to two dimensions; in
$d$ space dimensions we have instead \cite{Pamporova}
\begin{equation} \label{comp:lambda}
\lambda_{\rm b} = (4 \pi)^{d/2} \frac{1-d/4}{\Gamma(2-d/2)}
\frac{\epsilon_a^{1-d/2}}{m^{d/2}},
\end{equation}
with $\Gamma(x)$ the Gamma function. Note the absence of any
temperature-dependence in the effective theory (\ref{GP}). At zero
temperature, exactly the same effective theory was obtained by Hausmann,
using a self-consistent Green function method \cite{Haussmann}. (See
Ref.\ \cite{Pamporova} for a derivation along the lines presented here.)
This is, as that author argued, because in the BEC limit, the critical
temperature $T_0$ is much smaller than the dissociation temperature
$T_{\rm diss} \approx \epsilon_a$ at which the tightly bound fermion
pairs are broken up by thermal fluctuations. Hence, for all
temperatures in the range $T \leq T_0 << T_{\rm diss}$ we are
effectively in the zero-temperature regime \cite{Haussmann}.
The Gross-Pitaevski theory (\ref{GP}) describes a gapless, purely
propagating mode, viz.\ the Goldstone mode associated with the
spontaneously broken global U(1) symmetry---the so-called
Anderson-Bogoliubov mode \cite{bosepaper}.
The spacetime-dependent effective theory of a superconductor can only be
derived in a few special cases \cite{AT}: one being close to the
transition temperature where the expansion parameter is $1/T$ and
another being close to the absolute zero of temperature where the
expansion parameter is $1/|\bar{\Delta}|$. Outside these regimes, the
effective theory depends on the ratio $|\nabla|/\partial_0$, and an
expansion in both time derivatives and gradients is not possible. It is
amusing to note that in going from the weak-coupling BCS limit to the
strong-coupling BEC limit, we move from one valid regime to the other.
\section{Dual Theory \label{sec:dual}}
In this section we investigate the dual formulation of the
time-independent Ginzburg-Landau theory. This approach, in which the
magnetic vortices of a superconductor play the central role, originates
from lattice studies that started more than two decades ago
\cite{BMK,Peskin,TS,HeMu,DaHa,Kleinerttri,Bartholomew,Savit}. These
were in turn instigated by the success of the Kosterlitz-Thouless theory
describing the phase transition in a superfluid film as the unbinding of
vortex-antivortex pairs \cite{Berezinskii,KT73}. In the dual
formulation of the Ginzburg-Landau theory, the superconductor-to-normal
phase transition is understood as a proliferation of magnetic vortices.
A detailed presentation of these matters as well as an extensive list of
references to the literature can be found in Ref.\ \cite{GFCM}.
\subsection{Electric Current Loops}
To account for the magnetic interaction we couple the Ginzburg-Landau
theory in the usual, minimal way to a vector potential ${\bf A}$. We
also rescale the pair field such that $\beta$ times the Hamiltonian
becomes
\begin{equation} \label{gl:H}
{\cal H}_{\rm GL} = \left|(\nabla - 2 i e {\bf A})\phi\right|^2 +
m_\phi^2 |\phi|^2 + \lambda |\phi|^4 +
\frac{1}{2} (\nabla \times {\bf A})^2 + \frac{1}{2 \alpha} (\nabla \cdot
{\bf A})^2 ,
\end{equation}
where we added a gauge-fixing term with parameter $\alpha$. To acquire a
physical understanding of what this Hamiltonian describes \cite{habil}, we
recall that a $|\phi|^4$-theory gives a field theoretic description of
strings with contact repulsion \cite{Symanzik}. This equivalence rests on
Feynman's observation \cite{Feynman50} that the Green function
\begin{equation} \label{gl:start}
G({\bf x}) = \int_{\bf k} \frac{{\rm e}^{i {\bf k}
\cdot {\bf x}}}{{\bf k}^2 + m_\phi^2}
\end{equation}
can be expressed as a path integral. An easy way to see this is to
invoke Schwinger's proper-time method, which is based on Euler's form
\begin{equation} \label{gl:gamma}
\frac{1}{a^z} = \frac{1}{\Gamma(z)} \int_0^\infty \frac{\mbox{d} \tau}{\tau} \,
\tau^z \, {\rm e}^{- \tau a},
\end{equation}
to write the right-hand side of (\ref{gl:start}) as \cite{proptime}:
\begin{eqnarray} \label{gl:green}
G({\bf x}) &=& \int_0^{\infty} \mbox{d} \tau \, {\rm e}^{-\tau m_\phi^2}
\int_{\rm k} {\rm e}^{i {\bf k} \cdot {\bf x} } {\rm e}^{-\tau{\bf k}^2}
\nonumber \\ &=& \int_0^{\infty} \mbox{d} \tau \, {\rm e}^{-\tau m_\phi^2}
\left( \frac{1}{4 \pi \tau} \right)^{3/2} {\rm e}^{-\frac{1}{4} {\bf
x}^2/\tau}.
\end{eqnarray}
The factor $(1/4 \pi \tau)^{3/2} \exp(-\frac{1}{4} {\bf x}^2/\tau)$
appearing here can be interpreted as describing a Brownian string
trajectory, showing that one endpoint of the string (located at ${\bf
x}$) has a Gaussian distribution with respect to its other endpoint
(located at the origin) \cite{PathI}. If we imagine the string to be
composed of $N$ links, each of length $a$, then in the limit where $N
\rightarrow \infty$, $a \rightarrow 0$, the variable $\tau$
parameterizing the string stands for $\tau = N a^2/6$. The integration
over $\tau$ in (\ref{gl:green}) indicates that the string can be
arbitrary long, but the weighing factor $\exp(-m_\phi^2 \tau)$
exponentially suppresses long ones in the normal phase where $m_\phi^2 >
0$. To understand the physical meaning of this factor, let us return to
the discrete string model and write
\begin{equation} \label{tension}
m_\phi^2 = 6 [\sigma(a) -\sigma_{\rm cr}]/a,
\end{equation}
so that it becomes
\begin{equation}
{\rm e}^{-m_\phi^2 \tau} = {\rm e}^{-\sigma_{\rm eff}(a) L},
\end{equation}
with $\sigma_{\rm eff}(a) = \sigma(a) -\sigma_{\rm cr}$ the effective
string tension, and $L = N a$ the length of the string. The continuum
limit is obtained by simultaneously letting $a \rightarrow 0$ and
$\sigma(a) \rightarrow \sigma_{\rm cr}$, in such a way that the
right-hand side of (\ref{tension}) tends to the finite value $m_\phi^2$.
We thus see that the factor $\exp(-m_\phi^2 \tau)$ weighs strings
according to their lengths.
Following Feynman \cite{Feynman50}, we write the right-hand side of
(\ref{gl:green}) as a path-integral, i.e., as a sum over all possible string
trajectories having one endpoint at ${\bf x}(0)=0$ and the other at ${\bf
x}(\tau) = {\bf x}$ \cite{Feynman48}:
\begin{equation} \label{gl:feynrep}
G({\bf x}) = \int_0^{\infty} \mbox{d} \tau \int_{{\bf x}(0)=0}^{{\bf
x}(\tau)={\bf x}} \mbox{D} {\bf x}(\tau') \, {\rm e}^{-S_0},
\end{equation}
with the (Euclidean) action
\begin{equation} \label{gl:world}
S_0 = \int_0^\tau \mbox{d} \tau' \left[\tfrac{1}{4} \dot{\bf x}^2 (\tau') +
m_\phi^2 \right],
\end{equation}
where $\dot{\bf x}(\tau) = \mbox{d} {\bf x}(\tau)/\mbox{d} \tau$.
In a similar way, also the partition function of the free theory, which
written as a functional integral reads
\begin{equation}
Z_0 = \int \mbox{D} \phi^* \mbox{D} \phi \, \exp\left(- \int_{\bf x} {\cal H}_0
\right),
\end{equation}
can be be represented as a {\it path} integral, this time involving only
closed strings:
\begin{eqnarray}
\ln(Z_0) &=& -\ln [ {\rm Det} ({\bf p}^2 + m_\phi^2)] = - {\rm Tr} \ln (
{\bf p}^2 + m_\phi^2) \\
&=& \int_0^{\infty}
\frac{d\tau}{\tau} {\rm e}^{-\tau m^2_\phi} \int_{\bf k} {\rm
e}^{-\tau {\bf k}^2} = \int_0^\infty \frac{d \tau}{\tau} \oint \mbox{D} {\bf
x} (\tau') {\rm e}^{-S_0}. \nonumber
\end{eqnarray}
Here, we used Euler's form (\ref{gl:gamma}) in the limit of small $z$.
An additional factor $1/\tau$ arises because one can start traversing a
closed string anywhere along the loop. The $|\phi|^4$-interaction in
the Ginzburg-Landau model can be shown to result in the additional term
\cite{Parisi}
\begin{equation}
S_\lambda = - \lambda \int_0^{\tau} \mbox{d} \tau_l' \mbox{d} \tau_k' \, \delta \left[
{\bf x} (\tau_l') - {\bf x} (\tau_k') \right]
\end{equation}
in the action, which gives an extra weight each time two strings---one
parameterized by $\tau_l'$ and one by $\tau_k'$---intersect. Physically, it
represents a repulsive contact interaction between strings. Finally, the
coupling of the field $\phi$ to the magnetic vector potential ${\bf A}$ via
the electric current
\begin{equation}
{\bf j}_e = - 2 e i (\phi^* \nabla \phi - \phi \nabla \phi^*) -2 (2e)^2
{\bf A} |\phi|^2
\end{equation}
with a charge $2e$, results in the extra term \cite{GFCM}
\begin{equation}
S_e = 2ie \int_0^{\tau} \mbox{d} \tau' \, \dot{\bf x} (\tau') \cdot {\bf A}[{\bf
x}(\tau')],
\end{equation}
showing that the strings described by the Ginzburg-Landau theory carry an
electric current. Pasting the pieces together, we conclude that the
partition function of the Ginzburg-Landau theory can be equivalently
represented as a grand canonical ensemble of fluctuating electric current
loops, of arbitrary length and shape \cite{Copetal}:
\begin{equation} \label{eloops}
Z = \int \mbox{D} {\bf A} \, {\rm e}^{-\frac{1}{2} \int_{\bf x} \left[(\nabla
\times {\bf A})^2 + \frac{1}{\alpha} (\nabla \cdot {\bf A})^2 \right] }
\sum_{N=0}^{\infty} \frac{1}{N!} \prod_{l=1}^N \left[ \int_0^\infty
\frac{\mbox{d} \tau_l}{\tau_l} \oint \mbox{D} {\bf x}(\tau'_l) \right] {\rm e}^{-
S_{\rm GL}},
\end{equation}
with the action
\begin{eqnarray} \label{action}
S_{\rm GL} &=& \sum_{l=1}^N \int_0^{\tau_l} \mbox{d} \tau_l' \left\{\tfrac{1}{4}
\dot{\bf x}^2(\tau_l') + m_\phi^2 + 2ie \dot{\bf x}(\tau'_l) \cdot {\bf
A}[{\bf x}(\tau'_l)]\right\} \nonumber \\ && + \lambda \sum_{l,k=1}^N
\int_0^{\tau_l} \mbox{d} \tau_l' \int_0^{\tau_k} \mbox{d} \tau_k' \, \delta \left[
{\bf x} (\tau_l') - {\bf x} (\tau_k') \right] .
\end{eqnarray}
On entering the superconducting phase, characterized by a sign change in
the mass term of the Ginzburg-Landau theory, the effective string
tension approaches zero as
\begin{equation}
\sigma_{\rm eff}(a) \propto m_\phi^2 \propto (T-T_{\rm c})^{2 \nu},
\end{equation}
with $\nu$ the correlation length exponent, and the electric current loops
proliferate. This proliferation is also signaled by the absolute value
$|\phi|$ of the pair field which then develops a vacuum expectation value,
identifying it as the order parameter.
\subsection{Disorder Parameter}
The dual formulation of the Ginzburg-Landau gives a similar
representation of the partition function as the path integral
(\ref{eloops}); however this time not in terms of electric current, but
of magnetic vortex loops. To derive it, we start with a trick
\cite{fort} and introduce a (hypothetical) magnetic monopole at some
point ${\bf z}$ inside the superconductor. A monopole is a source of
magnetic flux. Due to the Meissner effect, the flux lines emanating
from the monopole are squeezed into a flux tube. In this way, we
managed to create a magnetic vortex at zero external field.
Electrodynamics in the presence of a monopole was first described by
Dirac \cite{Dirac} who argued that the physical local magnetic induction
${\bf h}$ is given by the combination $\nabla \times {\bf A}({\bf x}) -
{\bf B}^{\rm P}({\bf x})$. The subtracted plastic field
\begin{equation} \label{subtr}
B_i^{\rm P} ({\bf x}) = \Phi_0 \int_{L_{\bf z}} {\rm d} y_i \, \delta ({\bf
x} - {\bf y}),
\end{equation}
with $\Phi_0 = \pi/e$ the magnetic flux quantum, removes the field of the
so-called Dirac string running along some path $L_{\bf z}$ from the location
${\bf z}$ of the monopole to infinity. On account of Stokes' theorem, the
plastic field satisfies the equation
\begin{equation} \label{rhom}
\nabla \cdot {\bf B}^{\rm P} ({\bf x}) = - \rho_{\rm m} ({\bf x}),
\end{equation}
with $\rho_{\rm m} ({\bf x}) = \Phi_0 \, \delta ({\bf x} - {\bf z})$ the
monopole density.
We continue by writing $\phi({\bf x})$ as
\begin{equation} \label{bcs:London}
\phi({\bf x}) = \bar{\phi} \, {\rm e}^{2i \varphi ({\bf x})},
\end{equation}
where $\bar{\phi}$ is a constant solution of the mean-field equation. This
approximation, where the phase of the order field is allowed to vary in
space while the modulus is kept fixed, is called the London limit. The
Ginzburg-Landau Hamiltonian then becomes after integrating out the phase
field $\varphi$:
\begin{equation} \label{HP}
{\cal H}_{\rm GL}^{\rm P} = \frac{1}{2} (\nabla \times {\bf A} - {\bf
B}^{\rm P})^2 + \frac{1}{2} m_A^2 A_i \left( \delta_{i j} - \frac{\partial_i
\partial_j }{\nabla^2} \right) A_j + \frac{1}{2\alpha}(\nabla \cdot {\bf
A})^2,
\end{equation}
where
\begin{equation} \label{photonmass}
m_A = 2e |\bar{\phi}|
\end{equation}
is the inverse magnetic penetration depth $m_A = \lambda_{\rm L}^{-1}$. We
gave ${\cal H}$ the superscript ${\rm P}$ to indicate the presence of the
monopole. While the Dirac string is immaterial in the normal phase, Nambu
\cite{Nambu} argued that it acquires physical relevance in the
superconducting phase where it serves as the core of the magnetic vortex or
Abrikosov flux tube originating at the monopole.
The energy $E_V$ of this configuration is easily obtained by
substituting the solution of the field equation for the gauge field
\begin{equation} \label{clas}
A_i ({\bf x}) = \int_{\bf y} G_{i j} ({\bf x} - {\bf y})
\left[\nabla \times {\bf B}^{\rm P} ({\bf y})\right]_j,
\end{equation}
with $G_{i j}$ the gauge-field Green function
\begin{equation} \label{gfprop}
G{i j} ({\bf x}) = \int_{\bf k} \left(
\frac{\delta_{i j} - k_i k_j/{\bf k}^2}{{\bf k}^2+m_A^2} + \alpha
\frac{k_i k_j}{{\bf k}^4} \right) {\rm e}^{i {\bf k} \cdot {\bf x}},
\end{equation}
back into the Hamiltonian (\ref{HP}). The energy is divergent in the
ultraviolet because in the London limit, where the mass $|m_\phi|$ of
the $\phi$-field is taken to be infinite, the vortices are considered to
be ideal lines. For a finite mass, a vortex core has a typical width of
the order of the coherence length $\xi =1/|m_\phi|$. This mass
therefore provides a natural ultraviolet cutoff to the theory. Omitting
the irrelevant (diverging) monopole self-interaction, one finds
\cite{Nambu}
\begin{equation} \label{moncon}
E_V = \frac{1}{2} g^2 \int_{L_{\bf z}} {\rm d} x_i \int_{L_{\bf z}} {\rm d}
y_i\, G ({\bf x} - {\bf y})
= \sigma_V L_{\bf z} .
\end{equation}
Here,
\begin{equation} \label{Yuka}
G ( {\bf x} ) = \int_{\bf k} \frac{ {\rm e}^{i {\bf k} \cdot {\bf
x}}}{{\bf k}^2+m^2_A} = \frac{1}{4 \pi} \frac{{\rm e}^{-m_A |{\bf x}|}}{
|{\bf x}|}
\end{equation}
is the scalar Green function, $g$ abbreviates the combination
\begin{equation} \label{g}
g = \Phi_0 m_A,
\end{equation}
while $L_{\bf z}$ denotes the (infinite) length of the Dirac string and
\begin{equation} \label{M}
\sigma_V= \frac{1}{8\pi} g^2 \ln\left(
\frac{|m_\phi|^2}{m_A^2} \right) = \frac{1}{4\pi} g^2 \ln (
\kappa_{\rm GL} )
\end{equation}
its line tension, first calculated by Abrikosov \cite{Abrikosov}. The
value $\kappa_{\rm GL} = 1/\sqrt{2}$ of the Ginzburg-Landau parameter
\begin{equation} \label{gl:gl}
\kappa_{\rm GL}= \frac{|m_\phi|}{m_A} = \frac{\lambda_{\rm L}}{\xi},
\;\;\;\;\; \mbox{or} \;\;\;\;\; \kappa^2 _{\rm GL} =
\frac{\lambda}{2e^2},
\end{equation}
separates the type-II regime $(\kappa_{\rm GL} > 1/\sqrt{2})$, where
isolated vortices can exist, from the type-I regime $(\kappa_{\rm GL} <
1/\sqrt{2})$, where a partial penetration of an external field is
impossible. Remembering that $L_{\bf z}$ was the Dirac string, we see
from (\ref{moncon}) that in the superconducting phase it indeed becomes
the core of the magnetic vortex originating at the monopole, as was
first observed by Nambu \cite{Nambu}.
The operator $V(L_{\bf z})$ describing the monopole with its emerging
flux tube \cite{Marino,KRE} is easily constructed by noting that in the
functional-integral approach, a given field configuration is weighted
with a Boltzmann factor $\exp \left(-\int_{\bf x} {\cal H}_{\rm GL}^{\rm
P}\right)$, with the Hamiltonian given by Eq.\ (\ref{HP}). From this we
infer that the explicit form of the vortex operator is
\begin{equation} \label{dop}
V(L_{\bf z}) = \exp \left\{ \int_{\bf x} \left[(\nabla \times {\bf A})
\cdot {\bf B}^{\rm P} - \tfrac{1}{2} \left({\bf B}^{\rm P}
\right)^2 \right] \right\}.
\end{equation}
We next demonstrate that this operator can be used to distinguish the
superconducting from the normal phase \cite{Marino,KRE}. To this end,
let us consider the correlation function $\langle V( L_{{\bf z}} ) V^*(
L_{\bar {\bf z}}) \rangle$, where $V^*(L_{\bar {\bf z}})$ describes an
additional antimonopole brought into the system at ${\bar {\bf z}}$,
with $L_{\bar {\bf z}}$ being the accompanying Dirac string running from
infinity to ${\bar {\bf z}}$. Since all the integrals involved are
Gaussian, this expectation value can be evaluated directly. We proceed,
however, in an indirect way for reasons that will become clear when we
proceed, and first linearize the functional integral over the gauge
field by introducing an auxiliary field $\tilde{\bf h}$. In the gauge
$\nabla \cdot {\bf A} = 0$, which corresponds to setting $\alpha=0$ in
the Hamiltonian (\ref{HP}), we find \cite{MA}
\begin{eqnarray} \label{Vdu}
\lefteqn{\langle V( L_{\bf z} ) V^*(L_{\bar {\bf z}}) \rangle = }
\\ && \int {\rm D}
{\bf A} {\rm D} \tilde{\bf h} \exp \left\{ \int_{\bf x} \left[ -\frac{1}{2}
\tilde{\bf h}^2 + i \tilde{\bf h} \cdot (\nabla \times {\bf A} -
{\bf B}^{\rm P}) - \frac{m_A^2}{2} {\bf A}^2 \right] \right\},
\nonumber
\end{eqnarray}
where the plastic field satisfies (\ref{rhom}) with the monopole density
given by
\begin{equation} \label{mdens}
\rho_{\rm m} ({\bf x}) = \Phi_0 [ \delta ({\bf x}-{\bf z})- \delta({\bf
x} - {\bar {\bf z}})].
\end{equation}
To appreciate the physical relevance of the auxiliary field, let us
consider its field equation
\begin{equation}
\label{phys}
\tilde{\bf h} = i (\nabla \times {\bf A} - {\bf B}^{\rm P}) = i {\bf h}.
\end{equation}
It tells us that apart from a factor $i$, the auxiliary field
$\tilde{\bf h}$ can be thought of as representing the local magnetic
induction ${\bf h}$.
The integral over the vector potential is easily carried out by
substituting the field equation for ${\bf A}$,
\begin{equation}
\label{fieldeq}
{\bf A} = \frac{i}{m_A^2} \nabla \times \tilde{\bf h},
\end{equation}
back into (\ref{Vdu}), with the result
\begin{equation} \label{Vdua}
\langle V( L_{\bf z} ) V^*(L_{\bar {\bf z}}) \rangle =
\int {\rm D} \tilde{\bf h} \, \exp \left\{ -\frac{1}{2} \int_{\bf x} \left[
\frac{1}{m_A^2}(\nabla \times \tilde{\bf h})^2 + \tilde{\bf h}^2 \right]
- i \int_{\bf x} \tilde{\bf h} \cdot {\bf B}^{\rm P} \right\}.
\end{equation}
This shows that the magnetic vortex described by the plastic field ${\bf
B}^{\rm P}$ couples to the fluctuating massive vector field $\tilde{\bf
h}$, with a coupling constant $g$ introduced in Eq.\ (\ref{g}). As the
temperature approaches the critical temperature from below, $\bar{\phi}$
tends to zero, so that the vortex decouples from the auxiliary field
$\tilde{\bf h}$. The finite magnetic penetration depth in the
superconducting phase is reflected by the mass term in (\ref{Vdua}).
After carrying out the integral over $\tilde{\bf h}$ in (\ref{Vdua}), we
obtain for the correlation function
\begin{eqnarray} \label{vv*}
\lefteqn{\langle V( L_{\bf z} ) V^*(L_{\bar {\bf z}}) \rangle =} \\ &&
\exp \biggl\{ - \frac{1}{2} \int_{{\bf x}, {\bf y}} \bigl[
\rho_{\rm m} ({\bf x})\, G({\bf x} - {\bf y})
\, \rho_{\rm m} ({\bf y}) + m_A^2 B^{\rm P}_i
({\bf x}) \, G({\bf x} - {\bf y}) \, B^{\rm P}_i ({\bf y}) \bigr]
\biggr\}. \nonumber
\end{eqnarray}
The first term in the argument of the exponential function contains a
diverging monopole self-interaction for ${\bf x} = {\bf y}$. This
divergence is irrelevant and can be eliminated by defining a
renormalized operator
\begin{equation}
\label{renopV}
V_{\rm r} (L_{\bf z}) = V(L_{\bf z}) \exp\left[ \tfrac{1}{2} \Phi_0^2 G(0)
\right].
\end{equation}
The second term in the argument is the most important one for our purposes.
It represents a Biot-Savart interaction between two line elements ${\rm d} x_i$
and ${\rm d} y_i$ of the magnetic vortex (see Fig.~\ref{fig:biotsavart}).
\begin{figure}
\begin{center}
\epsfxsize=6cm
\mbox{\epsfbox{biotsavart.eps}}
\end{center}
\vspace{-.5cm}
\caption{Biot-Savart interaction (wiggly line) between two line elements
${\rm d} x_i$ and ${\rm d} y_i$ of a magnetic vortex (straight line).
\label{fig:biotsavart} }
\end{figure}
For the renormalized operators we find
\begin{equation} \label{correlation}
\langle V_{\rm r}( L_{\bf z} ) V_{\rm r}^*( L_{\bar {\bf z}} ) \rangle =
\exp(-\sigma_V L_{{\bf z}{\bar {\bf z}}} ) \exp\left(\frac{\Phi_0^2}{4 \pi}
\frac{ {\rm e}^{-m_A L_{{\bf z}{\bar {\bf z}}} } }{ L_{{{\bf z}}{\bar
{\bf z}}} } \right),
\end{equation}
where $L_{{\bf z}{\bar {\bf z}}}$ denotes the length of the flux tube
connecting the monopole at ${{\bf z}}$ with the antimonopole at ${\bar
{\bf z}}$. Initially, the two Dirac strings may run to infinity along
two arbitrary paths. Due to the string tension, however, they join on
the shortest path $L_{{{\bf z}}{\bar {\bf z}}}$ connecting the
monopoles.
The correlation function (\ref{correlation}) behaves differently in the
two phases. In the superconducting phase, where $m_A \neq 0$, the first
factor dominates. It represents the confining linear potential between
the monopole and the antimonopole. As a result, the correlation
function decays exponentially for distances larger than $1/\sigma_V$:
\begin{equation} \label{correlator}
\left\langle V_{\rm r}( L_{z{\bar z}} ) V_{\rm r}^*(L_{z{\bar z}})
\right\rangle \rightarrow 0.
\end{equation}
On the other hand, in the normal phase, where the gauge field is
massless, the confinement factor in the correlation function
(\ref{correlation}) disappears, while the argument of the second
exponential turns into a pure Coulomb potential. The correlation
function remains, consequently, finite for large distances:
\begin{equation}
\left\langle V_{\rm r}( L_{z{\bar z}} ) V_{\rm r}^*(L_{z{\bar z}})
\right\rangle \rightarrow 1,
\end{equation}
thus indicating a proliferation of magnetic vortices in the normal
phase. This demonstrates that the vortex operator can be used as an
order parameter to distinguish the two phases. Since it develops a
vacuum expectation value not in the superconducting, but in the normal
phase it is referred to as a {\it disorder} parameter.
It is interesting to consider in detail the limit $T \uparrow T_{\rm
c}$, where the coupling constant $g = \Phi_0 m_A$ tends to zero and the
magnetic vortex decouples from the massive vector field \cite{MA}. In
Eq.\ (\ref{Vdua}), this limit yields the constraint $\nabla \times
\tilde{\bf h} = 0$ which can be solved by setting $\tilde{\bf h} =
\nabla \gamma$. The correlation function then takes the simple form
\begin{equation} \label{dual:simple}
\langle V_{\rm r}({\bf z}) V_{\rm r}^*( {\bar {\bf z}}) \rangle = \int
{\rm D} \gamma \exp \left[ -\frac{1}{2} \int_{\bf x} (\nabla \gamma)^2 +
i \int_{\bf x} \gamma \rho_{\rm m} \right].
\end{equation}
In the absence of monopoles, the theory reduces to that of a free
gapless mode $\gamma$ that may be thought of as representing the
magnetic scalar potential:
\begin{equation}
\label{gammaid}
\nabla \gamma = i \nabla \times {\bf A} .
\end{equation}
This follows from combining the physical interpretation of the vector
field ${\bf h}$ (\ref{phys}) with the equation $\tilde{\bf h} = \nabla
\gamma$.
The correlation function (\ref{dual:simple}) can be put in the form
\begin{equation} \label{localrep}
\langle V_{\rm r}({\bf z} ) V_{\rm r}^*(
{\bar {\bf z}}) = \left\langle {\rm e}^{i \Phi_0 [\gamma ({\bf z})-
\gamma({\bar {\bf z}})] } \right\rangle,
\end{equation}
where the average at the right-hand side is taken with respect to the
free scalar theory. This equation shows that in the normal phase ($T >
T_{\rm c}$), the Dirac string looses its physical relevance, the
right-hand side depending only on the end points ${\bf z}$ and ${\bar
{\bf z}}$, not on the path $L_{{\bf z}{\bar {\bf z}}}$ connecting these
points. The notion of a magnetic vortex is of no relevance in this
phase because the vortices proliferate and carry no energy. This is the
reason for omitting any reference to vortex lines in the argument of $V$
in Eqs.\ (\ref{dual:simple}) and (\ref{localrep}).
\subsection{Magnetic Vortex Loops}
From the above results for a single vortex, we can now easily infer the
dual formulation of the Ginzburg-Landau theory. In this formulation,
the partition function is written as a grand canonical ensemble of
fluctuating magnetic vortex loops, of arbitrary length and shape:
\begin{equation} \label{pathZ}
Z = \int \mbox{D} \tilde{\bf h} \, {\rm e}^{-\frac{1}{2} \int_{\bf x} \left[
(\nabla \times \tilde{\bf h})^2/m_A^2 + \frac{1}{2} \tilde{\bf
h}^2 \right]} \sum_{N=0}^{\infty} \frac{1}{N!} \prod_{l=1}^N \left[
\int_0^\infty \frac{\mbox{d} \tau_l}{\tau_l} \oint \mbox{D} {\bf x}(\tau'_l) \right]
{\rm e}^{- S_{\rm dual}}
\end{equation}
with the dual action, [cf.\ Eq.\ (\ref{action})]
\begin{eqnarray}
S_{\rm dual} &=& \sum_{l=1}^N \int_0^{\tau_l} \mbox{d} \tau_l' \left\{\tfrac{1}{4}
\dot{\bf x}^2(\tau_l') + m_\psi^2 + i \Phi_0 \dot{\bf x}(\tau'_l) \cdot
\tilde{\bf h}[{\bf x}(\tau'_l)]\right\} \nonumber \\ && + u
\sum_{l,k=1}^N \int_0^{\tau_l} \mbox{d} \tau_l' \int_0^{\tau_k} \mbox{d} \tau_k'
\, \delta \left[ {\bf x} (\tau_l') - {\bf x} (\tau_k') \right] .
\end{eqnarray}
We have included here a mass term $m^2_\psi$ representing the intrinsic
vortex line tension [cf.\ Eq.\ (\ref{tension})], and also a contact
repulsion between the vortices parameterized by $u$.
The equivalent field representation of this dual theory reads
\cite{BS,Kaw,GFCM,KKR,MA,fort},
\begin{equation} \label{funcZ}
Z = \int \mbox{D} \tilde{\bf h} \mbox{D} \psi^{*} \mbox{D} \psi
\, \exp \left(- \int_{\bf x} {\cal H}_{\rm dual}\right)
\end{equation}
with the Hamiltonian
\begin{equation} \label{Hpsi}
{\cal H}_{\rm dual} = \frac{1}{2 m_A^2} (\nabla \times
\tilde{\bf h})^2 + \frac{1}{2} \tilde{\bf h}^2 + |(\nabla -i \Phi_0
\tilde{\bf h}) \psi|^2 + m_\psi^2 |\psi|^2 + u |\psi|^4 ,
\end{equation}
where the disorder field $\psi$ is minimally coupled to the massive
vector field $\tilde{\bf h}$, representing the local magnetic induction.
The dual formulation contains the same information as the original,
Ginzburg-Landau formulation. For example, the vortex line tension
(\ref{M}) appears in the dual theory as a one-loop on-shell mass
correction stemming from the graph depicted in Fig.\
\ref{fig:biotsavart} which we now interpret as a Feynman graph of the
dual theory (\ref{Hpsi}), with the straight and wiggly line denoting
respectively the $\psi$- and $\tilde{\bf h}$-field Green function. Also
the fixed points of the Ginzburg-Landau theory map onto those of the
dual field theory \cite{KKR}. And both formulations can be used to
study the critical behavior of the superconductor-to-normal phase
transition (see Refs.\
\cite{Peskin,Kleinerttri,GFCM,PRL,fort,Herbut,CaNo} for the dual
approach).
\\
\noindent
{\bf Acknowledgments}
\noindent
It is a pleasure to thank N. Antunes, L. Bettencourt, A. Leggett, D. Steer,
and G. Volovik for useful discussions during the NATO Winter School and
European Science Foundation (ESF) Workshop {\it Topological Defects and the
Non-Equilibrium Dynamics of Symmetry Breaking Phase Transitions}, Les
Houches, February 16-26, 1999, and M. Crisan for helpful correspondence.
This work is performed as part of a scientific network supported by the
ESF (see network's URL, http://www.physik.fu-berlin.de/$\sim$defect).
|
\section{Introduction}
The classical statistical theory of thermodynamical phenomena,
due largely to Boltzmann, Maxwell, and Gibbs, is one
of the cornerstones of 20th century physics. It describes equilibrium
phenomena ranging from gas dynamics over steam engines to crystals,
while its quantum extension accurately describes radiation phenomena,
metals, and superconductivity, to name but a few examples. Nature's
tendency to move towards equilibrium following a perturbation---captured by
Boltzmann's second law---implies that most every\-day-life phenomena are
indeed taking place in an equilibrated system, for which this theory is
applicable and eminently successful. For the brief {\em transitory} periods,
however, the time during which a system {\em approaches}
equilibrium, our bag of tricks---containing the tools of statistical
mechanics---is of little use. The canonical phenomena of this type are
relaxation or transport processes, phenomena which are usually termed
``irreversible'', and phase transitions for which the entropy is not a
constant.
The standard approach to deal with such situations is to study the
$N$-body dynamics of the system, with a Hamiltonian that includes an
interaction term (in equilibrium statistical mechanics the Hamiltonian
is a sum of non-interacting one-body terms) and the construction of
equations that follow the $N$-particle distribution function through
time: the Boltzmann equation (see, e.g., \cite{prigogine}).
This approach suffers from the drawback
that it can only be solved in perturbation theory, which obscures the
relation to the ``exact'' formalism of thermodynamics. In this paper,
we would like to explore the possibility that a formalism well-known from
engineering---Shannon's statistical theory of information---provides a
bridge between equilibrium and non-equilibrium statistical phenomena,
and that its quantum extension (developed primarily in support of the
recent efforts in quantum computation and communication) represents an
adequate framework to investigate certain quantum statistical phenomena that
have so far resisted a satisfying treatment. Naturally, however, we
should not expect that the classical and quantum theory of information
provides a complete theory of all non-equilibrium phenomena. For
most dynamics with complicated time-dependent interactions and
many-body correlations, a transport-equation approach will still be
the only tractable alternative.
Standard non-equilibrium phenomena are usually termed
``irreversible'', an adjective that captures a practical aspect---a
direction of time---which, however, we know not to be
fundamental. Rather, time-reversal invariance guarantees that all dynamics can,
in principle, be reversed as long as the participating degrees of
freedom can be controlled. Even though this is clearly not always possible in
practice, it may appear as an oversight that a
practical limitation seems to be at the origin of a theorem---the
second law of thermodynamics. Indeed, as irreversibility is only
practical, so must be the second law. If we were, then, able to devise
a formalism in which the second law is replaced by a {\em conservation
law} for entropy (and in which case the second law would appear as
a corollary) we may then be in possession of a formalism that can
quantitatively describe even the {\em approach} to equilibrium and
other non-equilibrium statistical phenomena. It is the purpose of this
paper to point out that this formalism exists in the form of the
classical theory of information, introduced by
Shannon~\cite{shannon}. Its extension to the quantum
regime (see, e.g.,~\cite{steane} and references therein)
is particularly interesting as it consistently describes quantum unitary
dynamics which dictates that the von Neumann entropy---the quantum
extension of the Shannon entropy---is a {\em constant}.
In the next section we begin by describing the classical
statistical theory of information in physical terms (as
opposed to the more engineering-oriented approach given in most
textbooks~\cite{textbooks}). We then apply it to two classical
non-equilibrium statistical processes---meas\-ure\-ment, and
equilibration of an ideal gas---to demonstrate the use of the
formalism in physics. In Section 3 we formulate the quantum theory
with special emphasis on those aspects that differ from the classical
theory, and discuss the EPR paradox as an illustration. We
present an application to black hole formation and
evaporation---a quintessential non-equilibrium scenario---in Section
4. We close with conclusions and comments in Section 5. Readers
familiar with the information-theoretic approach to classical and
quantum statistical phenomena may skip directly ahead to Section 4.
\section{Classical Theory}
The intimate
relation between information theory and statistical mechanics has been
pointed out earlier by Jaynes~\cite{jaynes} in order to {\em justify}
statistical mechanics via information theory. Here, we use information
theory to {\em extend} statistical mechanics to the non-equilibrium
regime.
The concept of entropy was introduced by Shannon with respect to {\em
random variables}. For a random variable $X$ that can take on values
$x_1,\cdots,x_N$ with probabilities $p_1,\cdots,p_N$ respectively, the
Shannon {\em uncertainty} (or entropy) is given by
\begin{equation}
H(X)=-\sum_{i=1}^Np_i\log p_i\;. \label{shanent}
\end{equation}
Instead of random variables,
however, we may imagine any physical system with enumerable degrees of
freedom and enumerable states $x_i$. As is well-known and we show
below, the Shannon entropy then represents the {\em physical} entropy
of the system. In fact, this concept of entropy can be
expanded to cover continuous variables, where it will suffer from the
same ambiguity (redefinition up to a constant) as standard
thermodynamical entropy.
For the moment, let us confine ourselves to discrete degrees of
freedom and imagine that any continuous variables are {\em coarse-grained}
(either by assuming appropriate boundary conditions, or else
artificially.)
The relation to Boltzmann-Gibbs entropy becomes manifest if we
consider not {\em general} probability distributions
$\left\{p_i\right\}$, but an equilibrium distribution where the $p_i$
are given by the Gibbs distribution:
\begin{equation}
p_i=\frac1Ze^{-E_i/kT}\;, \label{boltz}
\end{equation}
where $E_i$ is the {\em energy} of state $x_i$, and $p_i$ then
represents the {\em probability} of $X$ to take on energy $E_i$. Note
that this probability
is normalized by the partition function $Z=\sum_ie^{-E_i/kT}$. Inserting (\ref{boltz}) into
Eq.~(\ref{shanent}) produces
\begin{equation}
H = \frac{\langle E\rangle}{kT}+\log Z = \frac1{kT}(\langle E\rangle -F)\; \label{class}
\end{equation}
and confirms that the Shannon entropy is just the standard physical entropy
in statistical mechanics and thermodynamics when rescaled by
the Boltzmann constant $k$:
\begin{equation}
S = k H\;.
\end{equation}
Above, we defined the free energy $F=-kT\log Z$ in the usual manner.
Similarly, thermodynamical averages are obtained via
\begin{equation}
\langle A\rangle = \frac1Z\sum_{i=1}^N A_i e^{-E_i/kT}
\end{equation}
for an observable $A$ that takes on the value $A_i$ in state $x_i$.
Returning to random variables for a moment, imagine an additional
variable $Y$ that takes on states $y_1,\cdots,y_N$ with probabilities
$p_1'\cdots,p'_N$. We can then define the
conditional probability of finding $X$ in state $x_i$, {\em given}
that $Y$ is in state $j$
\begin{equation}
p_{i|j}=\frac{p_{ij}}{p'_j}\;,
\end{equation}
where $p_{ij}$ is the {\em joint} probability to find $X$ in state
$x_i$ and simultaneously $Y$ in state $y_j$.
This concept will allow us to quantify {\em correlations} between
degrees of freedom, a particularly important task in non-equilibrium
systems. Indeed, equilibrium can be {\em defined} as the state where ``all
`fast' things have happened and all the `slow' things
not''~\cite{feynman}, which implies that all non-permanent correlations
have vanished in equilibrium.
Armed with conditional probabilities, we can define the {\em conditional
entropy} of system $X$ {\em given} that $Y$ is in, say, state $y_j$,
i.e., the entropy of $X$ if we
are fully aware that $Y$ is in state $y_j$, or in other words,
the {\em remaining} entropy of $X$ if
$Y$ is held fixed in state $y_j$. Naturally,
this is defined as
\begin{equation} H(X|y_j) = -\sum_{i}p_{i|j}\log{p_{i|j}}\;.
\end{equation}
Also, the {\em average conditional entropy} of $X$ given $Y$ is in
{\em any} fixed state, or quite generally is {\em known}, is then
\begin{equation}
H(X|Y)=\langle H(X|y_j)\rangle = -\sum_{ij}p_{ij}\log p_{i|j}\;. \label{condent}
\end{equation}
The vertical bar in the expression $H(X|Y)$ denotes the conditional
nature of the entropy, and is usually read as ``X given Y'', or ``X
knowing Y''.
Armed with the conditional (or remaining) entropy, we can find a
measure for the amount of correlation between two systems. This is
just the ordinary entropy minus the remaining entropy if one of the
system's variables are known: the {\em shared} entropy (also called
correlation, or mutual, entropy)
\begin{equation}
H(X:Y) = H(X)-H(X|Y)\;.
\end{equation}
This is the central quantity introduced by Shannon: the mathematical
measure of {\em information}\footnote{The colon between $X$ and $Y$ is
customarily used to indicate a shared entropy, and reminds us that
correlation entropy is symmetric: $H(X:Y)=H(Y:X)$.}. The relation
between unconditional (also called ``marginal'') entropies such as
$H(X)$ or $H(Y)$, mutual, and conditional entropies
are best visualized by {\em Venn diagrams}. In Fig.~1, the area of
each circle represents an entropy, whereas the union of both circles
represents the joint entropy $H(XY)$.
\begin{figure}
\caption{Entropy Venn diagram for two random variables $X$ and $Y$.}
\label{fig1}
\vskip 0.5cm
\par
\centerline{\psfig{figure=fig1.ps,width=5cm,angle=-90}}
\vskip 0cm
\par
\end{figure}
It is straightforward to see that these quantities can be translated
into thermodynamics, by replacing the arbitrary probability
distributions by equilibrium ones. We can see immediately, however,
why they play no role in equilibrium thermodynamics. The probability
of system $X$ to take on energy $E_i$ if $Y$ has energy $E_j$ is
trivial: it is just given by $Z^{-1}e^{-E_i/kT}$ simply {\em because} $X$ and
$Y$ are in equilibrium. Thus, in equilibrium, $H(X|Y)=H(X)$, and
$H(X:Y)=0$. Away from equilibrium, conditional and mutual {\em
thermodynamical} entropies become crucial, as we now
see.
\subsection{Measurement}
We first treat the dynamics of classical {\em measurement}. A
measurement involves the contact between two equilibrated systems,
usually at different temperatures. The measurement device is
constructed in such a manner as to induce correlations between some of
its variables---the ``pointer''---and the measured system's degrees of
freedom (those which we desire to measure). After the initial contact
between the systems and subsequent relaxation, equilibrium is
re-established but thermodynamics seems to offer a paradox: the
entropy of the measured system appears to have been
reduced. Furthermore, this reduced entropy {\em can} be used to
perform work---in apparent violation of the second law (this puzzle is
usually termed the {\em Maxwell demon} paradox, see, e.g., \cite{demon}).
While this dynamics is again practically irreversible, we can describe
what happens in terms of the entropies introduced above.
Before the measurement, the system (denoted by $S$) is independent of
the measurement device (denoted by $M$, see Fig.2a). They do not share
any entropy, which implies that knowledge of any one of the systems
will not allow any predictions about the other. Bringing the two
systems into contact introduces correlations, and reduces the {\em
conditional} entropy of both $S$ and $M$. Note that before
measurement, $H(S|M)\equiv H(S)$. The amount by which the conditional
entropy is reduced is of course just the {\em acquired information},
or shared entropy $H(S:M)$ (see Fig.~2b). This shared entropy plays a
fundamental thermodynamical role: for example it can be shown that
erasing it requires the dissipation of an equal amount of
heat~\cite{landauer}. Needless to say, the marginal entropy did not
really decrease in this process, but rather {\em stayed constant}.
In contrast, the conditional entropy of $S$ is reduced, as can be seen
by inspection of the diagram in Fig.~2b,
\begin{figure}[t]
\caption{Rearrangement of entropies in the measurement process. (a)
System $S$ and device $M$ are uncorrelated ($H(S:M)=0$). (b) Device
and system share entropy $H(S:M)$ and the conditional entropy of
both system and device are reduced.}
\label{fig2}
\vskip 0.5cm
\par
\centerline{\psfig{figure=fig2.ps,width=10cm,angle=-90}}
\vskip 0.5cm
\par
\end{figure}
\begin{equation}
H(S)\longrightarrow H(S|M) = H(S) - H(S:M)\;. \label{master}
\end{equation}
Turning Eq.~(\ref{master}) around:
\begin{equation}
H(S)=H(S|M)+H(S:M)
\end{equation}
demonstrates that non-equilibrium dynamics affects only the
distribution of $H(S)$ into either (conditional) entropy or
information, that the two however always add up to $H(S)$.
\subsection{Equilibration}
Another example of irreversible dynamics is the notorious
``perfume bottle'' experiment, in which a diffusive substance (let's
say, an ideal gas) is allowed to escape from a small container into a
larger one. Both the initial and the final state of the system is in
equilibrium; common wisdom however states that the entropy of the gas is {\em
increasing} during the process, reflecting the non-equilibrium
dynamics. We shall now show that this is not the case, by describing
the gas in the smaller container by a set of variables
$A_1,\cdots,A_n$, one for each molecule. The entropy $H(A_i)$ thus
represents the entropy per molecule. The entire
volume, on the other hand, is described by the {\em joint entropy}
\begin{equation}
H_{\rm gas}=H(A_1\cdots A_n)\;, \label{joint}
\end{equation}
which can be much smaller than the sum of per-particle entropies, the
standard (equilibrium) thermodynamical entropy $S_{eq}$
\begin{equation}
H(A_1\cdots A_n)\ll \sum_{i=1}^n H(A_i)=S_{eq}\;.
\end{equation}
The difference is given by the $n$-body correlation entropy
\begin{equation}
H_{\rm corr}= \sum_{i=1}^n H(A_i)-H(A_1\cdots A_n) \label{corr}
\end{equation}
which can be calculated in principle, but becomes cumbersome already
for more than three particles.
We see that in this description the molecules after occupying the
larger volume cannot be independent of
each other, as their locations are {\em in principle} correlated (as they
all used to occupy a smaller volume, see
Fig.~3a). These correlations are not manifest in two-- or even
three-body correlations, but are complicated $n$-body correlations
which imply that their positions are not independent, but linked by
the fact that they share initial conditions. Again, this state of
affairs can be summarized by turning around Eq.~(\ref{corr})
\begin{equation}
H(A_1\cdots A_n) =\sum_{i=1}^n H(A_i)-H_{\rm corr}\;.
\end{equation}
We assume that before the molecules are allowed to escape, they
are uncorrelated with respect to each other: $H_{\rm corr}=0$, and all
the entropy is given by the extensive sum of the per-molecule entropies.
After expansion into the larger volume, the standard entropy increases
because of the increase in available phase space, but this increase is
balanced by an increase in the correlation entropy $H_{\rm corr}$ in
such a manner that the actual joint entropy of the
gas, $H_{\rm gas}$, remains unchanged.
Note that this description is not, strictly speaking, a {\rm redefinition} of
thermodynamical entropy. While in the standard theory entropy is an
{\em extensive}, i.e., additive quantity for uncorrelated systems, the
concept of a thermodynamical entropy in the absence of equilibrium
distributions has been formulated as the number of ways to realize a
given set of occupation numbers of states of the joint system
(which gives rise to
(\ref{shanent}) by use of Stirling's approximation, see, e.g.,
\cite{wannier}) and is thus fundamentally {\em non-extensive}.
Assuming the systems $A_i$ are uncorrelated reduces $H(A_1\cdots A_n)$
to the extensive sum
$\sum_{i=1}^{n}H(A_i)$, and thus to an entropy proportional to the volume
the systems inhabit. From a calculational point of view the present
formalism does not represent a great advantage in this case, as the correlation
entropy $H_{\rm corr}$ can only be obtained in special situations,
when only few-body correlations are important.
The examples of non-equilibrium processes treated here (measurement
and equilibration) suggest that:
\begin{quote}
\em In a thermodynamical equilibrium or non-equilibrium process, the
unconditional (joint) entropy of a closed system remains a constant.
\end{quote}
This formulation of the second law directly reflects probability
conservation (in the sense of the Liouville theorem),
and allows a quantitative description of the amount by
which the conditional entropy is decreased in a measurement, or the
amount of per-particle entropy is increased in an equilibration process.
\begin{figure}
\caption{Diffusion of an ideal gas from a small into a larger container. (a)
The molecules with entropy $H(A_1\cdots A_n)$ occupy the smaller
volume, and their correlation entropy is zero. (b) The molecules
have escaped into the larger container, which increases the sum of
the per-particle entropies and increases the correlation entropy
commensurately such that the overall entropy remains unchanged.}
\label{fig3}
\vskip 0.5cm
\par
\centerline{\psfig{figure=fig3.ps,width=9cm,angle=-90}}
\vskip 0cm
\par
\end{figure}
\section{Quantum Theory}
As the classical non-equilibrium mechanics described above is founded
on the classical theory of information, its quantum
extension is built on the quantum theory of information introduced
recently~\cite{schum,ca1,ca2}.
\subsection{Equilibrium}
For our purposes, equilibrium quantum statistical mechanics can be
summarized in a few equations. For a system described by
Hamiltonian\footnote{In the following, $H$ stands for the Hamiltonian,
while entropies are denoted by the symbol $S$.}
$H$ and partition function (we set $\beta=1/kT$ from now on)
\begin{equation}
Z={\rm Tr}\, e^{-\beta H}\;,
\end{equation}
the density matrix can be written as
\begin{equation}
\varrho = \frac{e^{-\beta H}}Z \label{equilib}
\end{equation}
while the free energy is
\begin{equation} F = -\frac1\beta \log Z\;.
\end{equation}
Accordingly,
\begin{equation}
\log \varrho = \beta F -\beta H \label{logrho}
\end{equation}
and, defining the internal energy $U={\rm Tr}\,\varrho H$, we obtain the
equivalent of Eq.~(\ref{class})
\begin{equation}
S = \beta (U-F)
\end{equation}
where
\begin{equation}
S(\varrho)=-{\rm Tr}\, \varrho \log \varrho\label{vnentropy}
\end{equation}
is the quantum entropy of the state described by the density
matrix $\varrho$, introduced by von Neumann~\cite{vn}.
While we used equilibrium
expressions to motivate (\ref{vnentropy}), it is in fact valid even
when an equilibrium expression such as (\ref{equilib}) does not
exist. Just as the classical entropy (\ref{joint}), this entropy
remains a constant under {\em any} dynamics, reversible or
irreversible. This is in fact more obvious in the quantum
case, as the density matrix $\varrho$ is known to evolve in a unitary
manner
\begin{equation}
\varrho(t)=U(t)\varrho(0)U^\dagger(t) \label{unitary}
\end{equation}
which immediately implies, using (\ref{vnentropy}) and the cyclic
property of the trace, that
\begin{equation}
\frac{\rm d}{{\rm d}t} S(t) = 0\;.
\end{equation}
Inserting (\ref{equilib}) into (\ref{vnentropy}) on the other
hand allows us to recover the
Boltzmann-Gibbs-Shannon entropy (\ref{shanent}), with the
probabilities given by
\begin{equation}
p_i=\frac1Z \,e^{-\beta E_i}
\end{equation}
with $E_i$ the eigenvalues of $H$. In general, when considering the
diagonal elements of $\varrho$ in a basis distinct from the eigenbasis
of $H$, the von Neumann entropy is a lower bound on the
Boltzmann-Gibbs-Shannon entropy
\begin{equation}
S(\varrho)\leq -\sum_ip_i\log p_i\;,
\end{equation}
where the equality holds for density matrices $\varrho$ that are
diagonal, in which case quantum statistical mechanics is formally
identical to the classical description. Differences arise for
non-diagonal $\varrho$. The off-diagonal terms signal the presence of
quantum {\em superpositions} and the potential for {\em
entanglement}---a form of ``super-correlation''.
As we shall see, entanglement requires a radical departure from the
classical description, and an extension of the above formalism to a
non-equilibrium quantum statistical mechanics.
\subsection{Non-equilibrium}
As mentioned earlier, in classical mechanics equilibrium between two
ensembles $A$ and $B$ implies that all ``fast'' degrees of freedom are
independent (no correlations) whereas the ``slow'' degrees are
considered to be static. This is usually achieved by waiting for times
larger than the relaxation time.
The situation is dramatically
different in quantum mechanics. As we shall see, entanglement
introduces a type of super-correlation that cannot be undone by
letting the system equilibrate, not even if the two systems are
separated by space-like distances.
As an example, consider the joint system $AB$ where $A$ and $B$ are
half-integral spin states with eigenstates $|\uparrow\rangle$ and
$|\downarrow\rangle$. It is then possible to construct a wavefunction for
the joint system $AB$ which makes it mathematically and logically
impossible to attribute a {\em state} to either $A$ or $B$ by itself: the
well-known EPR state
\begin{equation}
|\Psi_{AB}\rangle=\frac1{\sqrt{2}}\left(|\uparrow\ua\rangle-|\downarrow\da\rangle\right)\;. \label{epr}
\end{equation}
However, both $A$ and $B$ can be described by {\em reduced} density
matrices, obtained by tracing $B$ or $A$ out of
the joint matrix $\varrho_{AB}$
\begin{equation}
\varrho_{A(B)}={\rm Tr}_{B(A)}\varrho_{AB}=
\frac12\biggl(|\uparrow\rangle\langle\uparrow|+|\downarrow\rangle\langle\downarrow|\biggr)\;,
\end{equation}
where ${\rm Tr}_{B(A)}$ denotes the partial trace over $B(A)$.
As these density matrices are diagonal, the quantum entropy is just
equal to the classical one
\begin{equation}
S(A)=S(B)=1
\end{equation}
if we agree to take base-2 logarithms and count entropy in
``bits''. The joint entropy $S(AB)$ on the other hand is {\em not}
equal to 2, i.e., the entropy is {\em non-extensive}. As we mentioned
earlier, this implies that correlations are present and calls for a
non-equilibrium formalism. Things are worse here.
For this wavefunction, the quantum entropy {\em vanishes} (it is a
pure state: the only non-vanishing eigenvalue of the density matrix
$\varrho_{AB}= |\Psi_{AB}\rangle \langle \Psi_{AB}|$ is 1.) This well-known property of
quantum mechanically entangled systems is known as the {\em
non-monotonicity} of quantum entropy (see, e.g., \cite{wehrl}) and
forces us to rethink the equilibrium formalism that we recapitulated
earlier. We will proceed in a manner similar to the non-equilibrium
classical mechanics of the previous section, by introducing quantum
{\em conditional} and {\em mutual} entropies. As in the classical
case, the conditional quantum
entropy then would reveal to us the entropy of a quantum system {\em
given} we know the state of another system it is entangled with,
while the quantum mutual entropy would reflect the amount of
correlation between the systems. In contrast to the classical
situation, quantum conditional entropies can be {\em negative}, while
the mutual entropy can {\em exceed} the classically allowed limit (hence the
term super-correlation.) This formalism has turned out to be useful in
the information-theoretic analysis of quantum
measurement~\cite{ca2,meas}, as well as the description of the
non-equilibrium physics of quantum information
transmission~\cite{channel}.
Guided by the classical case, we are tempted to define the conditional
quantum entropy of system $A$ given the state of $B$ by
\begin{equation}
S(A|B)= S(AB)-S(B)\;,
\end{equation}
i.e., the quantum entropy of the joint system minus the entropy of $B$
(as that is given). This structure then suggests an expression for the
{\em conditional amplitude matrix} $\varrho_{A|B}$, which we need to
formulate the non-equilibrium dynamics. This matrix, first introduced
in \cite{ca1}, is a well-defined Hermitian operator on the joint
Hilbert space of $A$ and $B$ (see \cite{CAG}) defined by
\begin{equation}
\varrho_{A|B}=\exp[\log\varrho_{AB}-\log({\bf 1}_A\otimes\varrho_B)]
\label{condmat}
\end{equation}
which allows us to write
\begin{equation}
S(A|B) = -{\rm Tr} \varrho_{AB}\log \varrho_{A|B}
\end{equation}
in analogy with (\ref{condent}). In contrast to the classical
conditional probability $p_{i|j}$, the conditional amplitude matrix
can have eigenvalues {\em exceeding} unity, which reflect the quantum
inseparability of the system.
The mutual quantum entropy can be defined in an analogous manner
\begin{equation}
S(A:B) = S(A)-S(A|B)
\end{equation}
as the marginal (unconditional) quantum entropy of $A$ minus the
``remaining'' entropy $S(A|B)$. Consequently, we can extend the useful
Venn diagram technique (Fig.~1) to the quantum regime, and just
replace $H$ by $S$ (Fig.~\ref{fig4}a).
The peculiarity of quantum superpositions such as
the EPR wavefunction Eq.~(\ref{epr}) is immediately apparent in its
Venn diagram (Fig.~\ref{fig4}b).
\begin{figure}[t]
\caption{Quantum entropy Venn diagrams. (a) Definition of
joint [$S(AB)$] (the total area), marginal [$S(A)$ or $S(B)$],
conditional [$S(A|B)$ or $S(B|A)$] and mutual [$S(A:B)]$ entropies for a
quantum system $AB$ separated into two subsystems $A$ and $B$; (b)
their respective values for the EPR pair.
\label{fig4} }
\vskip 0.25cm
\centerline{\psfig{figure=fig_quantum.ps,width=3.75in,angle=0}}
\vskip 0.5cm
\end{figure}
More generally, a mixed state $\varrho=\sum_i p_i |i\rangle\langle i|$ can
always be ``purified'', i.e., written as the partial trace over a pure
state $|\psi\rangle=\sum_i\sqrt(p_i)|i\rangle|i\rangle$ by means of the Schmidt
decomposition, while being represented by a Venn diagram such as
Fig.~\ref{fig4}b but with entries $\{-S,2S,-S\}$ instead of
$\{-1,2,-1\}$, where $S=-\sum_i p_i\log p_i$. Furthermore, the diagram
technique and the use of quantum entropies can easily be extended to
understand the quantum correlations between three systems. An
instructive example is the description of the EPR
paradox~\cite{reality}, which we briefly summarize as it is relevant
to the discussion of black holes which follows.
Imagine a wavefunction such as (\ref{epr}), with the particles in
question separated by space-like distances. Imagine further that at
each of these separated locations, measurements of the spin-projection
are performed in either the $x$ or the $z$ direction. Beyond the
quantum bipartite system described by Eq.~(\ref{epr}), which we denote by
$Q_1Q_2$ in the following, we introduce Hilbert spaces for the
measurement devices, the ``ancillae'' $A_1$ and $A_2$ rigged to measure
the polarization of $Q_1$ and $Q_2$ respectively (see Fig.~\ref{meas}).
\begin{figure}[t]
\caption{Measurement of EPR pair $Q_1Q_2$ by devices $A_1$ and $A_2$.
\label{meas} }
\vskip 0.25cm
\centerline{\psfig{figure=epr.ps,width=2.0in,angle=-90}}
\end{figure}
Depending on whether same (Fig.~\ref{epr1}) or orthogonal
(Fig.~\ref{epr2}) polarizations are measured at
the remote locations, the measurement devices are either correlated or
independent. However, in both cases, the entanglement between quantum
systems and measurement devices is more complicated, and even in case
the measurement devices appear uncorrelated (Fig.~\ref{epr2}b), subtle
entanglement persists.
\begin{figure}[h]
\caption{ (a) Quantum entropy diagram for the EPR measurement of same
spin-projections: e.g., $A_1$ and $A_2$ both measure $\sigma_z$. (b)
Reduced diagram obtained by tracing over the quantum states $Q_1$ and
$Q_2$ (the dashed line surrounds degrees of freedom
traced out, i.e., averaged over) reflecting the correlation between
the measurement devices.
\label{epr1} }
\vskip 0.5cm
\centerline{\psfig{figure=epr1.ps,width=4.0in,angle=-90}}
\end{figure}
\begin{figure}
\caption{ (a) Quantum entropy diagram for the EPR measurement of orthogonal
spin-projections, e.g., $A_1$ measures $\sigma_Z$ while $A_2$ records
$\sigma_x$. (b) Reduced diagram as above. In this case the measurement
devices show zero correlation, while entanglement persists between
quantum system and measurement devices.
\label{epr2} }
\vskip 0.5cm
\centerline{\psfig{figure=epr2.ps,width=4.0in,angle=-90}}
\vskip 0.5cm
\end{figure}
\par
\section{Black hole Formation and Evaporation}
The discovery of Hawking radiation~\cite{hawking} appears
to have plunged quantum mechanics into a deep crisis, as it seems to
imply that the evaporation of black holes violates unitarity (for a
review, see, e.g.,~\cite{preskill}). Below, we formulate the
``information-loss'' problem in terms of the formalism described here,
and argue for a consistent description in terms of quantum
non-equilibrium thermodynamics.
\subsection{Black hole entropy and information paradox}
Black holes have the remarkable property that they are fully described
by very few variables---a non-rotating non-charged black hole by only
one, its mass. Bekenstein~\cite{bekenstein} and Hawking~\cite{hawking}
determined that an {\em entropy} can be defined for a Schwarzschild
black hole which
is given entirely in terms of the area $A$ inside the event horizon
\begin{equation}
S_{BH}=\frac14 A\;.
\end{equation}
This area, in turn, is just $A=4\pi R^2$ where $R$ is the radius of
the black hole given (in units where $\hbar=G=1$) by $R=2M$, so that
the black hole entropy is specified entirely in terms of the black
hole mass $M$
\begin{equation}
S_{BH}=4\pi M^2\;.
\end{equation}
While a number of reasonings lead to this expression, including the counting
of microscopic quantum states that give rise to a black hole,
Hawking~\cite{hawking1}
pointed out that the process of thermal evaporation of a black hole
leads to an ``information paradox''. If we assume that the black hole
is formed from a quantum mechanically pure state $S=0$, the entropy of
the purely thermal
blackbody radiation left behind {\em after evaporation} should be of
the order $\sim M^2$, i.e., a pure state evolved to a
mixed one. This contradicts the unitary evolution of quantum states
Eq.~(\ref{unitary}), according to which (as we have pointed out
repeatedly) the entropy of a closed system is a constant, in this
particular case the constant zero.
Several avenues have been proposed to escape this conclusion, and we
will focus here on the most conservative explanation, namely that
Hawking radiation is effectively {\em non-thermal} (in the sense that
quantum correlations between the radiation and the state of the black
hole exist in principle), and that a pure state {\em is} formed after
evaporation, only that it is impossible to distinguish it from
purity~\cite{page,thooft,ds}. We first note that beyond the
information paradox pointed out by Hawking, as observed by
Zurek~\cite{zurek} we also need to match the
black hole entropy $S_{BH}$ with the entropy of approximately thermal
radiation $S_{\rm rad}\sim T^3_{H}$ with black hole temperature
$T_H=(8\pi M)^{-1}$. We then proceed to propose a scenario in which
this might be achieved.
\subsection{Black hole formation from a pure state}
Of course, black holes do not form by the ``collapse'' of a pure
state. Rather, we can imagine that part of a pure state
with marginal entropy $S_{\rm rad}\equiv \Sigma$ disappears behind an event
horizon. Let us divide space just before the collapse into a region
{\it PBH}\ (the proto-black-hole) and $R$, the remainder. As the entire
system is pure ($S=0$), we know that $S_{\rm rad}=S_{PBH}$. The
entropy diagram for this situation can be constructed as described in
the previous section, and is shown in Fig.~\ref{fig5}a.
\begin{figure}
\caption{Venn diagrams for black hole formation. (a) Just before
collapse. (b) After collapse. $\Sigma$ denotes the entropy of the
proto-black-hole, while $S_{BH}$ is the Bekenstein-Hawking entropy,
and $\Delta S$ is the entropy deficit.}
\label{fig5}
\vskip 0.5cm
\par
\centerline{\psfig{figure=fig5.ps,width=12cm,angle=-90}}
\vskip 0cm
\par
\end{figure}
The degrees of freedom in $R$ are
practically inaccessible after the collapse of
the region {\it PBH}, but we should keep in mind that they are
{\em entangled} with {\it PBH}\ in such a manner that the entire system,
($R,{\it PBH}$),
is pure. In the language of quantum information theory, $R$ is a
``reference'' system that ``purifies'' {\it PBH}. The gravitational
collapse of region {\it PBH}\ forms an intriguing problem. While we can
assume the radiation inside it to be purely thermal, with energy
$E\sim T^4$ and corresponding entropy $\Sigma\sim4/3\,T^3$,
the entropy of the {\em collapsed} state is $S_{BH}=4\pi M^2$, lower
than $\Sigma$. In fact, it was shown by Zurek~\cite{zurek} that the
entropy ${\rm d} S$ accreted by a black hole (which we can take to be of the
radiation type) is larger than the corresponding entropy increase of the black
hole itself
\begin{equation}
{\rm d} S \approx 4/3 \,\,{\rm d} S_{BH}\;,
\end{equation}
and the same mismatch occurs in the evaporation process.
In statistical physics this is not an alarming state of affairs, but
rather is the usual scenario in a non-equilibrium phase
transition. Here, we shall mask our ignorance about the dynamics which
produces the black hole out of radiation by assigning a new {\em
phase} to the black hole matter, and discuss the process in which
the radiation with entropy $\Sigma$ {\em condenses} to a phase with
entropy $S_{BH}$.
During the condensation from the proto-black-hole state to the
black-hole ({\it BH}) state, excess
entropy $\Delta S$ has to be radiated away ($T_{H}\Delta S$ is the
equivalent of the latent heat in a first-order phase transition) . While we
cannot offer a detailed picture of this transition, we assume that
this radiation is emitted just outside the forming horizon, and
represents the bremsstrahlung of the accelerated particles
accreting on the black hole. This gives rise, then, to the system
depicted in Fig.~4b, where the bremsstrahlung $R'$ is entangled
with both $R$ and the black hole {\it BH}, with marginal entropy
$S(R')=\Delta S
= \Sigma - S_{BH}$. During the phase transition,
the entropy of the {\it PBH}\ system remains constant, but is distributed
over the joint system ({\it BH},$R'$):
\begin{equation}
\Sigma =S(PBH)=S(R',BH)=S(BH)+S(R'|BH)=S_{BH}+\Delta S\;.
\end{equation}
The ``missing'' entropy $\Delta S$ therefore is contained in radiation
$R'$ emitted during the collapse.
This scenario, which is the time-reverse of the evaporation
process considered next, naturally leads to a radiation field $R'$
that is causally uncoupled from the black hole, as
$S(BH:R')=0$. Tracing over the ``reference'' field $R$ leads to the
trivial entropy diagram diagram $\{S_{BH},0,\Delta S\}$. We need to
keep in mind, however, that just as in the EPR situation described
previously, the wavefunctions of $R'$ and the black hole are linked
via entanglement with the quantum degrees of freedom $R$.
\subsection{Evaporation of black holes}
The processes of black hole formation and evaporation can be
considered time-reverse images of each other.
Evaporation of black holes occurs through the formation of
virtual particle--anti-particle pairs of energy $2dE$ close to the
horizon due to quantum mechanical tunneling in the
strong gravitational field. If one of the members of the pair
disappears behind the horizon while the other manages to escape, the
escaping particle appears to have a black-body spectrum with
temperature $T_{H}$, while the energy of the black hole is reduced by
$dE$. The paradox occurring here thus appears to be the same as the one
encountered in the condensation process. How does the radiation pick
up the extra entropy? In terms of quantum information theory, the creation of a
particle--anti-particle pair is akin to
the creation of an EPR state with vanishing entropy,
described by the entropy diagram in Fig.~\ref{fig4}b. However,
just as in standard first-order ``evaporation'' transitions, the
black hole has to provide in addition the latent heat for ``decondensation'',
i.e., the energy to create
the entropy $\Delta S$. Thus, a pair created with
$2{\rm d} E$ and temperature
$T_{H}$ will not reduce the black hole mass by an amount $\d E$, but by
\begin{equation}
\Delta E={\rm d} E-T_{H}\Delta S\;,
\end{equation}
which restores the entropy and energy balance.
The entropy of the escaping particle is ${\rm d} S\sim T_{H}^3$
while at the same time the entropy of the black hole is reduced by
\begin{equation}
{\rm d} S_{BH}=4\pi\left(M^2-(M-\Delta E)^2\right) =
\frac{{\rm d} E}{T_H}-\Delta S\;.
\end{equation}
Arguments have been raised (see the reviews~\cite{preskill} and in
particular~\cite{susskind}) that seem to imply that information stored
in correlations and entanglement between the black hole and its
surrounding radiation field cannot be retrieved, even in principle.
These arguments rest on the assumption that the (low-energy) quantum
fields live in a Hilbert space that is of the product form ${\mathbf
H}_{\rm in}\otimes{\mathbf H}_{\rm out}$, and an application of the
quantum no-cloning theorem. While the fields do live in a product
Hilbert space, the wavefunction of an EPR pair
created at the event horizon of the black hole indirectly becomes
entangled with the hole the moment one of the particles crosses the horizon
(even though the quantum fields are separated by space-like distances)
and the combined quantum state becomes inseparable. This situation is
not unlike the scenario we noted in the formation of the black hole,
where the accreted particle and the radiation it emits when tumbling
into the black hole can be considered an entangled, EPR-type state
(albeit with real rather than virtual energy). Just as in that case
the radiation $R'$ shared no entropy with the black hole, neither does
the Hawking radiation, while still being entangled with it. Thus, the
Hawking radiation carries ``information'' about the inside of the hole
in the same manner as the measurement of EPR partners separated by
space-like distances reveals correlations in measurement devices that
are at space-like distances.
Yet, a fundamental problem remains
that is unlikely to be solved within the present formalism. The
Hawking radiation---while emitted in a unitary manner and while
information loss certainly does not take place---remains causally
uncorrelated to the black hole as long as the horizon separates the
black hole entropy from the radiation field. In a sense, we have to
wait until the last moment---the disappearance of the black hole---for
the entropy balance to be restored. This appears to put a severe
strain on current black hole models, as it is hard to imagine that
this much entropy can be stored in an ever-shrinking black hole. This
problem is likely due to our incomplete understanding of late-stage
black holes, rather than a problem intrinsic to quantum mechanics.
An alternative solution would present itself if the Bekenstein-Hawking
entropy could be understood in terms of a {\em conditional} entropy.
In that case, entropy flow from the black hole to the outside via the
formation of virtual pairs is understood easily, as the member of the
pair that crosses the horizon not only has negative energy but also
negative conditional entropy (see Fig.~\ref{fig4}b). As a
conditional entropy can become as negative as the marginal entropy of
the system it is a part of, we can circumvent the argument that ``the
black hole cannot store the information until the end because it runs
out of quantum states'', because the radiation could ``borrow'' as
much entropy as necessary from the black hole
until the horizon has
disappeared. Within the present framework, there appears to be no
physical picture which would suggest that the Bekenstein-Hawking
entropy is in fact conditional. It is not inconceivable, however, that
a quantum statistical information theory extended to curved space-time
would reveal such a state of affairs.
\section{Conclusions}
We have used a formalism developed in the exploration of quantum
com\-pu\-ters---quantum information theory---to describe quantum processes away
from thermodynamical equilibrium, such as the formation and
evaporation of black holes. The formalism emphasizes the {\em
conservation} of entropy, and is particularly useful in situations
where entropy is distributed over two or three systems. We emphasize
that great care is needed in using the concepts of entropy and
information consistently: information, for example, can {\em never} be
``stored'' in one system (e.g., a black hole). Rather, information is
a measure of correlation {\em between} two systems, which implies that
information is {\em always} stored in correlations. The analysis of
information storage in black hole formation and evaporation presented
here is a simple application of these rules to a scenario in which
black holes are considered special states of matter with an equation
of state different from that of radiation (or usual matter).
Transitions between those states occur continuously as the specific
heat of black hole matter is negative~\cite{hawking}. As a
consequence, radiation and black hole matter are unstable at any time,
and transitions must occur as long as matter of either kind is
present. Yet, a consistent formulation of the correlations between
radiation and matter shows that entropy is not created during the
process, and consequently that information is conserved. Still, the mechanism
by which the pure state is restored in the last stages of black hole
evaporation may require deeper insights into quantum gravitational
dynamics, and possibly an extension of information theory to curved
space-time.
\noindent{\bf Acknowledgments}
\noindent We are indebted to H. A. Bethe for many useful discussions,
in particular for suggesting to us to address the
impact of negative entropies on quantum statistical mechanics. This
work was supported in part by NSF Grants PHY 94-12818 and PHY
94-20470, and by a grant from DARPA/ARO through the QUIC Program
(\#DAAH04-96-1-3086). N.J.C. is Collaborateur Scientifique of the
Belgian National Fund for Scientific Research.
|
\section{INTRODUCTION}
In today's standard inflationary scenario the amplified vacuum
fluctuations
of the inflaton field provide the seeds for large-scale structure
formation
\cite{linde90,padmanabhan93,brandenberger92/liddle93}. It is thus of
interest to understand the effect of these quantum
fluctuations on the universe dynamics and the implications they may have
on cosmological observations.
In most inflationary models (exponential inflation), the geometry of
spacetime during the period of inflation can be reasonably well described
by
de Sitter spacetime. This spacetime also seems favored by quantum
cosmology since most
accepted ``initial condition'' proposals, which are either quantum
``tunneling from nothing'' \cite{vilenkin84/86} or ``no-boundary''
condition \cite{hartle83}, both predict it.
It has been recently pointed out that fluctuations in the stress-energy
tensor of quantum fields may be important for some states in curved
spacetimes or even flat spacetimes with non-trivial topology \cite{kuo93}.
Hu and Phillips, for instance, have computed the energy density
fluctuations
of quantum states in spatially closed Friedmann-Robertson-Walker models
and have
shown that these could be as important as the energy density itself \cite
{hu97}. If so, these fluctuations may induce relevant back-reaction
effects
on the gravitational field (the spacetime geometry).
In this paper we compute the fluctuations of the stress-energy
tensor for a scalar field in de Sitter spacetime. We consider a massless
minimally coupled field in the Euclidean vacuum state, which is a de
Sitter invariant state. The motivation for considering the massless
minimal coupling case
is the fact that it mimics the behavior of gravitons in a curved
background
as well as that of the perturbations of the inflaton field in usual
inflationary models. As for the state, the reasons why we chose the
Euclidean vacuum are twofold. On the one hand, the high degree of symmetry
makes the computations simpler. On the other hand, there are at least two
serious physical motivations. First, it naturally arises in exponential
inflation models and for a massive field it has been shown to be the state
to which any other state tends asymptotically in a de Sitter background
\cite{hawking83}; and second, it is selected by the most popular boundary
conditions in quantum cosmological models \cite{halliwell89}.
Here we compute the two-point correlation for spacelike separated points
of the
stress-energy tensor and find that these fluctuations are important.
Therefore the back reaction of these fluctuations on the spacetime
geometry
could be relevant, and will be the subject of further work within the
context of stochastic semiclassical gravity and Einstein-Langevin
equation \cite{martin98}.
Related to this fact it is worth mentioning that Abramo, Brandenberger and
Mukhanov \cite{abramo97a,abramo97b} have shown that the second order
contribution
to back reaction of inflaton and gravitational perturbations during the
inflationary period can be important even below the ``self-reproduction''
scale.
Although the spirit of our approach to the back-reaction problem is
slightly
different, we believe a partial connection with their work may exist.
These issues will also be addressed in future investigations.
The plan of the paper is the following. In section \ref{sec2} we give a
very brief review of de Sitter spacetime, its properties and the solution
of the Klein-Gordon equation in such a background. In section \ref{sec3},
following the proposal made in \cite{garriga93}, we discuss the invariant
vacuum state for a massless minimally coupled field that we are going to
use. In section \ref{sec4} we deal with the fluctuations of the
stress-energy tensor of the scalar field in de Sitter background and
compute the noise kernel (i.e. the two-point correlations), which
characterizes these fluctuations, for spacelike separated points. We
discuss our results in the final section.
Throughout the paper we will use the $(+,+,+)$ sign convention of Misner,
Thorne and Wheeler's book \cite{misner71}.
\section{DE SITTER SPACETIME} \label{sec2}
In this section we give a brief summary of some useful properties and
definitions related to de Sitter spacetime. For a more detailed
exposition, the reader is referred to \cite{hawking73/mottola85,allen85}.
Four dimensional de Sitter space-time has positive constant curvature and
is thus maximally symmetric. Its ten dimensional group of isometries is
$O(4,1)$. It can be represented as a hyperboloid embedded in a five
dimensional spacetime with Minkowskian metric $\eta _{AB}$, so that
$\eta_{AB}\xi ^A(x)\xi ^B(x)=H^{-2}$, where the Hubble constant $H$ is
related to the scalar curvature $R=12H^2$ and $\xi ^A(x)$ is the position
vector in the
five-dimensional Minkowskian embedding spacetime (with $A,B=0,\ldots ,4$)
corresponding to the point $x$ of de Sitter. One can define the biscalar
\begin{equation}
\label{00}Z(x,y)\equiv H^2\eta _{AB}\xi ^A(x)\xi ^B(y) ,
\end{equation}
$Z>1$ for timelike separated points, $Z=1$ for points connected by null
geodesics and $Z<1$ for spacelike separated points. It is worth
emphasizing that there is no geodesic connecting two points with
$Z<-1$. For points which are geodesically connected an alternative
expression to (\ref{00}) is \cite{allen85}:
\begin{equation}
\label{0}Z(x,y)=\cos \sqrt{\frac{R\,\sigma (x,y)}6},
\end{equation}
where $\sigma (x,y)\equiv \frac 12 s^2(x,y)$ with $s(x,y)$ being the
geodesic distance
between the two points. We will use the closed coordinates $(\eta ,\chi
,\theta ,\varphi )$, which cover the whole de Sitter spacetime. The line
element reads:
\begin{equation}
\label{01}ds^2=(H\sin \eta )^{-2}\left( -d\eta ^2+d\chi ^2+\sin ^2\chi
\left( d\theta ^2+\sin ^2\theta \,d\varphi ^2\right) \right),
\end{equation}
where $\eta \in (0,\pi )$ and $(\chi ,\theta ,\varphi )$ are the usual
hyperspherical parametrization of the $S^3$ spatial surfaces, which are
invariant under the $O(4)$ subgroup of isometries.
A scalar field of mass $m$ satisfies the Klein-Gordon equation
$\left( \Box -m^2-\xi R\right) \phi (x)=0$, where $\xi$ is a
dimensionless constant which determines the coupling of the field to the
spacetime curvature. We look for a set of mode solutions that can be
written as $u_{klm}(x)=H\sin \eta\,X_k(\eta )\,Y_{klm}(\chi ,\theta
,\varphi)$, where $Y_{klm}$ (with $k=0,1,2,\ldots ;l=0,1,\ldots ,k-1$ and
$m=-l,\ldots ,l$) are the $S^3$ spherical harmonics obeying
$\Delta^{(3)}Y_{klm}=-k(k+2)Y_{klm}$, which constitute a $(k+1)^2$
dimensional representation of $O(4)$. Substituting into the Klein-Gordon
equation, $X_k(\eta )$ satisfy the following equation:
\begin{equation}
\label{04}\frac{d^2X_k}{d\eta ^2}+\left\{ (k+1)^2+(H\sin \eta )^{-2}\left[
m^2+\left( \xi -\frac 16\right) R\right] \right\} X_k=0.
\end{equation}
The general solution to this equation is
\begin{equation}
\label{05}X_k(\eta )=(\sin \eta )^{-\frac 12}\left[ A_kP_{k+\frac 12%
}^\lambda (-\cos \eta )+B_kQ_{k+\frac 12}^\lambda (-\cos \eta )\right],
\end{equation}
where $\lambda \equiv \sqrt{\frac 94-\frac{12}R(m^2+\xi R)}$, the
coefficients $A_k$ and $B_k$ satisfy the normalization condition
$A_kB_k^{*}-A_k^{*}B_k=i\Gamma (k+\frac 32%
-\lambda )/\Gamma (k+\frac 32+\lambda )$ and $P_{k+\frac 12}^\lambda
(z)$ and $Q_{k+\frac 12}^\lambda (z)$ are associated Legendre functions of
the first
and second kind respectively. Given such a complete set of orthonormal
solutions, one can expand the field operator as $\hat{\phi }%
(x)=\sum_{klm}\left( \hat{a}_{klm}u_{klm}(x)+\hat{a}%
_{klm}^{+}u_{klm}^{*}(x)\right) $ and build the associated Fock space of
states in the usual way. If we require the vacuum state to be $O(4,1)$
invariant and of Hadamard type, $A_k$ and $B_k^{}$ are uniquely determined
\cite{allen87,chernikov68/schomblond76}:
\begin{eqnarray}
\label{07}
A_k&=&\frac \pi 4\frac{\Gamma (k+\frac 32-\lambda )}{\Gamma (k+\frac 32
+\lambda)}, \\
B_k&=&-\frac{2i}\pi A_k .
\end{eqnarray}
The corresponding state is the so-called Euclidean vacuum, also known as
Bunch-Davies vacuum.
\section{DISCUSSION ABOUT THE CHOSEN STATE} \label{sec3}
The most natural state to choose is the Euclidean vacuum, which is
invariant
under the de Sitter group of isometries and is a Hadamard state. This is
the
one selected in quantum cosmology when Hartle-Hawking no-boundary proposal
or Vilenkin's tunneling wave function are taken as initial conditions
\cite
{halliwell89}. In addition, when the field is massive, the field state
asymptotically tends to it \cite{hawking83}.
Both the Green's functions and the renormalized expectation value of the
stress-energy tensor have been computed for the Euclidean vacuum in
several
works \cite{bunch78,birrell84}:
\begin{equation}
\label{08}G^{\left( 1\right) }(x,y)\equiv \left\langle 0\right| \{
\hat{\phi }(x),\hat{\phi }(y)\} \left| 0\right\rangle
=\frac{%
2H^2}{(4\pi )^2}\,\Gamma\!\left( \frac 32-\lambda \right) \Gamma\!\left(
\frac 32+\lambda \right) F\!\left( \frac 32-\lambda ,\frac 32+\lambda
,2;\frac{1+Z(x,y)}2\right),
\end{equation}
\begin{eqnarray}
\label{09}
\left\langle
\hat{T}_{\mu \nu }(x)\right\rangle _{ren}=-\frac{g_{\mu \nu}(x)}{64\pi^2}
\left\{ m^2\left[ m^2+\left( \xi -\frac 16\right) R\right] \left[ \psi\!
\left( \frac 32-\lambda \right) +\psi\! \left( \frac 32+\lambda \right)
+\ln
\frac R{12m^2}\right] \right. \nonumber \\
\left. -m^2\left( \xi -\frac 16\right) R-%
\frac 1{18}m^2R-\frac 12\left( \xi -\frac 16\right)
^2R^2+\frac{R^2}{2160}%
\right\},
\end{eqnarray}
where $F(a,b,c;z)$ is a hypergeometric function and $\psi (z)\equiv
(d/dz)\,\ln \Gamma (z)$.
From (\ref{08}) one can see that the symmetrized two-point function $%
G^{\left( 1\right) }(x,y)$ has an infrared divergence (as $m^2\rightarrow
0$%
\/) in the minimal coupling case. On the other hand, the expectation value
of the stress-energy tensor remains finite, but is ambiguous, i.e., it
depends on the way the limit $m^2\rightarrow 0$, $\xi \rightarrow 0$ is
taken. In fact, Allen \cite{allen85} proved that in the massless minimally
coupled case there can be no Fock space built on a de Sitter invariant
vacuum. Allen and
Folacci suggested that in this case one should consider instead $O(4)$
invariant states \cite{allen87}. On the other hand Garriga and Kirsten
pointed out that such a peculiar behavior comes from the zero mode, which
corresponds to a constant solution of the Klein-Gordon equation that
appears
when $k=0$ in (\ref{04}) \cite{garriga93}. They also showed that a special
treatment of the zero mode
seems to give an invariant vacuum. The Hilbert space of states is then the
tensor product of the zero mode part, which is equivalent to the Hilbert
space for a one dimensional non-relativistic free particle, times the Fock
space corresponding to the non-zero modes (the whole space is not a Fock
space, in agreement with Allen's result). The field operator is written
\begin{equation}
\label{001}\hat{\phi }(x)=\sum_{%
{{klm} \atop {k\neq 0}}
}\left( \hat{a}_{klm}u_{klm}(x)+\hat{a}_{klm}^{+}u_{klm}^{*}(x)%
\right) +\frac H{\sqrt{2}\pi }\left( \hat{Q\,}\frac
12+\hat{P}\left(
\eta -\frac 12\sin 2\eta -\frac \pi 2\right) \right) ,
\end{equation}
where $\hat{Q}$ and $\hat{P}$ are the position and momentum
operators of the free particle Hilbert space and $\hat{a}_{klm}^{+}$
and $\hat{a}_{klm}$ are the creation and annihilation operators of the
Fock space associated to all the modes with $k\neq 0$.
Note that the two functions that multiply the operators $\hat{Q}$ and
$\hat{P}$, when divided by $\sin \eta$, are solutions of equation
(\ref{04}) for $k=0$ (with $m=\xi=0$). In fact, the constant
function multiplying the operator $\hat{Q}$ is precisely the aforesaid
zero mode. These operators satisfy the usual commutation relations:
\begin{equation}
\label{002}
\begin{array}{c}
\left[ \hat{a}_{klm},\hat{a}_{k^{\prime }l^{\prime }m^{\prime }}\right]
=0,\qquad \left[ \hat{a}_{klm},\hat{a}_{k^{\prime }l^{\prime}
m^{\prime }}^{+}\right] =\delta _{kk^{\prime }}\delta _{ll^{\prime}}
\delta_{mm^{\prime }},\qquad [ \hat{Q},\hat{P} ] =i, \\
\protect{} [ \hat{a}_{klm},\hat{Q} ] =0,\qquad [ \hat{a}_{klm},\hat{P} ]
=0 .
\end{array}
\end{equation}
The vacuum is defined as follows:
\begin{equation}
\label{003}
\begin{array}{c}
\hat{P}\left| 0\right\rangle =0 ,\\ \hat{a}_{klm}\left|
0\right\rangle =0\;,\,k\neq 0 .
\end{array}
\end{equation}
As a consequence of the first condition, this vacuum is not normalizable,
so strictly speaking it is not an
element of the Hilbert space but just a limit. However, as it has been
pointed out by Garriga and Kirsten, this fact is not so
strange: it also happens for the ground state of a non-relativistic free
particle. These authors also computed the renormalized expectation
value of the stress-energy tensor for this state:
\begin{equation}
\label{2}\left\langle \hat{T}_{\mu \nu }(x)\right\rangle
_{ren}=\frac{119}{138240\pi ^2}\,R^2\,g_{\mu \nu }(x)=\frac{119}{960\pi
^2}\,H^4\,g_{\mu \nu }(x).
\end{equation}
Two aspects should be stressed. First, the expression is de Sitter
invariant. Second, the energy density is lower than that of the $O(4)$
invariant
states introduced by Allen an Folacci, the stress-energy tensor of which
is not de Sitter invariant. It is also remarkable that the expectation
value for other states of observables such as the dispersion of the
smeared field or the stress-energy tensor tend
asymptotically ($\eta \rightarrow \pi $ or equivalently $t\rightarrow
\infty $, where $t$ is the proper time for a comoving observer) to the
same value as that of the invariant vacuum \cite{garriga93}.
\section{COMPUTATION OF THE STRESS-ENERGY TENSOR FLUCTUATIONS}
\label{sec4}
The quantity that characterizes the stress tensor fluctuations and which
determines, via the Einstein-Langevin equation, the fluctuations in the
spacetime geometry \cite{martin98} is the noise kernel, defined by the
symmetrized two-point correlation function $\frac 12\left\langle
\{\hat{t}_{\mu \nu}(x),\hat{t}_{\rho \sigma }(y)\} \right\rangle = \Re
\left\langle
\hat{t}_{\mu \nu}(x)\hat{t}_{\rho \sigma }(y) \right\rangle$,
where $\hat{t}_{\mu \nu}
(x)\equiv \hat{T}_{\mu \nu }(x)-\left\langle \hat{T}_{\mu \nu
}(x)\right\rangle $ and $\hat{T}_{\mu \nu }(x)\equiv\lim_{x^{\prime
}\rightarrow x}{\cal D}_{\mu \nu}\,(x,x^{\prime})\left( \hat{\phi
}(x)\hat{\phi
}(x^{\prime })\right) $ with ${\cal D}_{\mu \nu }(x,x^{\prime })\equiv
\left(\nabla_\mu ^x\nabla _\nu ^{x^{\prime}}-\frac 12g_{\mu \nu }(x)\nabla
_\alpha ^x\nabla ^{x^{\prime }\alpha }\right)$ (we have chosen
point-splitting regularization for convenience). It is easy to see that
$\left\langle
\hat{t}_{\mu \nu}(x)\hat{t}_{\rho \sigma }(y) \right\rangle$
is equivalent to $\left\langle \hat{T}_{\mu \nu }(x)\hat{T}_{\rho
\sigma }(y)\right\rangle -\left\langle \hat{T}_{\mu \nu
}(x)\right\rangle \left\langle \hat{T}_{\rho \sigma }(y)\right\rangle$.
Note that this expression is finite in the following sense: one can
compute it suitably regularized, then the potentially divergent terms
cancel and one can remove the regularization (letting $x^{\prime
}\rightarrow x$ and $y^{\prime }\rightarrow y$).
\begin{eqnarray}
\label{3}
\left\langle
\hat{T}_{\mu \nu }(x)\hat{T}_{\rho \sigma }(y)\right\rangle
-\left\langle \hat{T}_{\mu \nu }(x)\right\rangle \left\langle
\hat{T}_{\rho \sigma }(y)\right\rangle =\lim_{{x^{\prime }\rightarrow
x}\atop{y^{\prime }\rightarrow y}}{\cal D}_{\mu \nu }(x,x^{\prime}){\cal
D}_{\rho\sigma}(y,y^{\prime})&&\left( \left\langle \hat{\phi}(x)\hat{\phi}
(x^{\prime })\hat{\phi }(y)\hat{\phi }(y^{\prime })\right\rangle \right.
\nonumber \\
&&\left. -\left\langle \hat{\phi }(x)\hat{\phi }(x^{\prime
})\right\rangle \left\langle \hat{\phi }(y)\hat{\phi }(y^{\prime
})\right\rangle \right).
\end{eqnarray}
The expression inside the parenthesis is computed in the appendix, where
we find that it is finite as we pointed out above, and equals
$G^{+}(x,y)G^{+}(x^{\prime},y^{\prime})+G^{+}(x,y^{\prime})G^{+}
(x^{\prime},y)$. The final result after removing the regularization is:
\begin{eqnarray}
\label{6}
\left\langle
\hat{t}_{\mu \nu }(x)\hat{t}_{\rho \sigma }(y)\right\rangle &=&
\Bigl[
\nabla _\rho ^y\nabla _\mu ^xG^{+}(x,y)\nabla _\sigma ^y\nabla _\nu
^xG^{+}(x,y)+\nabla _\sigma ^y\nabla _\mu ^xG^{+}(x,y)\nabla _\rho
^y\nabla
_\nu ^xG^{+}(x,y) \nonumber \\
&& -g_{\mu \nu }(x)\nabla _\rho
^y\nabla _\alpha ^xG^{+}(x,y)\nabla _\sigma ^y\nabla ^{x\alpha
}G^{+}(x,y)-g_{\rho \sigma }(y)\nabla _\alpha ^y\nabla _\mu
^xG^{+}(x,y)\nabla ^{y\alpha }\nabla _\nu ^xG^{+}(x,y) \nonumber \\
&& +\frac 12g_{\mu \nu }(x)g_{\rho \sigma }(y)\nabla _\beta ^y\nabla
_\alpha ^xG^{+}(x,y)\nabla ^{y\beta }\nabla ^{x\alpha }G^{+}(x,y)\Bigr].
\end{eqnarray}
Thus, we need to determine the Wightman function $G^{+}(x,y)\equiv
\left\langle 0\right| \hat{\phi }(x) \hat{\phi }(y)\left|
0\right\rangle $ for our chosen state. If we consider the Euclidean vacuum
for a massive minimally coupled field and take the massless limit, the
Wightman function diverges. As already explained, this is connected to
the impossibility of having a Fock space built on a de Sitter invariant
vacuum in the massless case. To apply Garriga and Kirsten's special
treatment of the zero mode, we have to consider separately the
contribution
from the zero mode and that from all the rest. To compute the latter, one
can use the expression for the massive case, with the mass acting as a
regulator of the infrared divergence, subtract the contribution from the
zero mode, which contains all the infrared divergences, and then remove
the regulator, i.e., take the massless limit.
When $x$ and $y$ are spacelike separated, $G^{+}(x,y)=\frac 12\left(
G^{\left( 1\right) }(x,y)+G(x,y)\right) =\frac 12 G^{\left( 1\right)
}(x,y)= \Re G^{+}(x,y)$
since the commutator $G(x,y)\equiv \left\langle 0\right| [ \hat{%
\phi }(x),\hat{\phi }(y)] \left| 0\right\rangle $ vanishes for
causally disconnected points. In this case the noise kernel coincides with
expression (\ref{6}). We can also take advantage of the results in refs.
\cite{garriga93,allen87} to write:
\begin{equation}
\label{7}G^{\left( 1\right) }(x,y)=G_{NZM}^{\left( 1\right)
}(x,y)+\frac{H^2%
}{8\pi ^2}\left\langle 0\right| \hat{Q}^2\left| 0\right\rangle ,
\end{equation}
where
\begin{equation}
\label{8}G_{NZM}^{\left( 1\right) }(x,y)\equiv\frac R{48\pi ^2}\left[
\frac 1{1-Z(x,y)}-\ln \left( 1-Z(x,y)\right) -\ln (4\sin \eta _x\sin \eta
_y)-\sin^2\eta _x-\sin ^2\eta _y\right] .
\end{equation}
Several remarks are in order. $G_{NZM}^{\left( 1\right) }(x,y)$
corresponds
to the non-zero mode contribution whereas the contribution from the zero
mode reduces to $H^2/8\pi ^2 \,\left\langle 0\right| \hat{Q}%
^2\left| 0\right\rangle $ since $\hat{P}\left| 0\right\rangle =0$.
This
term is actually divergent but is independent of $x$ and $y$, so that it
will give no contribution to (\ref{6}). Furthermore, the term $\ln
(2\sin \eta _x)+\ln (2\sin \eta _y)+\sin ^2\eta _x+\sin ^2\eta _y$
will not contribute either, because in (\ref{6}) there are always
derivatives with respect to both $x$ and $y$ acting on $G^{+}(x,y)$.
Consequently, we only need to take into account the following part of the
two-point function $G^{(1)}(x,y)$:
\begin{equation}
\label{9}{\cal G}(x,y)\equiv \frac R{48\pi ^2}\left[ \frac 1{1-Z(x,y)}-\ln
\left( 1-Z(x,y)\right) \right] .
\end{equation}
If we consider spacelike separated points which are geodesically
connected,
we can use expression (\ref{0}) to derive the following results:
\begin{eqnarray}
\label{10}
\nabla _\mu ^x\sigma (x,y)=s(x,y)s_\mu (x), \\
\label{11}
\nabla _\rho ^y\nabla _\mu ^x\sigma (x,y)=s_\rho (y)s_\mu (x),
\end{eqnarray}
where $s^\mu (x)$ is the unit vector tangent at point $x$ to the geodesic
joining $x$ and $y$. The derivatives of $Z(x,y)$ are then
\begin{eqnarray}
\label{12}
\nabla _\mu ^xZ(x,y)=
\sqrt{\frac R{12}\left( 1-Z^2(x,y)\right) }\;s_\mu (x), \\ \nabla _\rho
^y\nabla _\mu ^xZ(x,y)=\frac R{12}\left( 1-Z^2(x,y)\right) s_\rho (y)
s_\mu (x).
\end{eqnarray}
Thus, after a few algebraic manipulations, we have
\begin{equation}
\label{13}
\nabla _\rho ^y\nabla _\mu ^xG^{+}(x,y)=
\nabla _\rho ^y\nabla _\mu ^x{\cal G}(x,y)=\frac{H^4}{4\pi ^2}
\frac{1+4Z(x,y)}{\left( 1-Z(x,y)\right) ^2}\,s_\rho (y)s_\mu (x) .
\end{equation}
Substituting this result into (\ref{6}), we get our final result for the
noise kernel corresponding to spacelike separated points:
\begin{eqnarray}
\label{14}
\left\langle
\hat{t}_{\mu \nu }(x)\hat{t}_{\rho \sigma }(y)\right\rangle
=\frac{H^8}{16\pi ^4}\left( \frac{1+4Z(x,y)}{\left( 1-Z(x,y)\right)
^2}\right)
^2 && \bigl[ 2 s_\mu (x)s_\nu (x)\,s_\rho (y)\,s_\sigma (y)+g_{\mu \nu
}(x)\,s_\rho (y)\,s_\sigma (y) \nonumber \\
&& +g_{\rho \sigma }(y)s_\mu
(x)s_\nu (x)+\frac 12g_{\mu \nu }(x)g_{\rho \sigma }(y)\bigr] .
\end{eqnarray}
\section{CONCLUSION}
Comparing (\ref{14}) with the ``square'' of (\ref{2}) one realizes that
the
contribution from the stress-energy fluctuations is at least as important
as that coming from the expectation value. Note that in order to compare
both
expressions it is convenient to consider the metric and the tangent vector
components as referred to an orthonormal base.
Of course, given that $H^4 << m_p^2 H^2$, this stochastic back-reaction
source is still much smaller than the dominant term which
drives inflation (usually coming from the potential of the inflaton scalar
field) and which is responsible for the near de Sitter geometry of the
spacetime. It seems interesting to point out that the
expectation value (\ref{2}) yields a negative value for the energy density
because of two facts. First, Ford and collaborators (see for instance ref.
\cite{kuo93}) have suggested that one may expect important stress-energy
fluctuations especially in those cases where the energy density is
negative. Second, this is in agreement with some of the results found in
\cite{abramo97b}.
It is only the contribution to the variations of the potential coming from
fluctuations of the inflaton field which are usually considered when
addressing the generation of large-scale gravitational inhomogeneities.
Expressions of the sort $\left\langle \phi ^2(x)\right\rangle
=\frac{H^3}{4\pi ^2}\Delta t$ or related ones \cite{ford82/linde82},
where proper infrared (related to initial conditions) and ultraviolet
cut-offs (only scales larger than the horizon are considered:
$\lambda_{phys}>H^{-1}$) have been imposed, are of frequent use in the
literature \cite{linde90,starobinsky82}.
These fluctuations for the smeared inflaton field can be interpreted as if
the smeared field were undergoing a sort of ``Brownian motion''
\cite{vilenkin83}. Such an
interpretation still applies to Garriga and Kirsten's vacuum as they show
when analyzing the dispersion of the smeared field \cite{garriga93}.
The fluctuations associated to a given cosmic scale arise from the
stochastic contribution, when leaving the horizon, from the modes of the
inflaton field perturbations with a wavelength corresponding to that
scale. The usual treatment just deals with first order variations of the
potential, which are assumed to provide the dominant contribution, and
does not take into account the back-reaction contribution from the kinetic
terms of the quantum perturbations of the inflaton field \cite{linde90}.
The latter can be modeled, at least partially, by a free massless
minimally coupled quantum scalar field. This is precisely the case we
considered in this paper. Furthermore, as mentioned above, we have found
that the contribution from the fluctuations of the stress-energy tensor
can be as important as the expectation value itself. Within the
inflationary context, this contribution to back reaction becomes
especially relevant above the ``self-reproduction'' scale and could have
important implications for the stochastic inflation approach, where this
effect is never taken into account.
One can try to extract information about large-scale fluctuations from
correlations
for spacelike separated points which are beyond the horizon distance \cite
{hawking82}. Unfortunately, for such points $Z(x,y)<-1$ and there is no
geodesic connecting them, thus one cannot directly use our results, which
were obtained using (\ref{10}) and (\ref{11}), which in turn rely on
the particular expression (\ref{0}) for geodesically connected points. If
one is interested in the $Z(x,y)<-1$ case, then the general expression
(\ref{00}) should be used. The relative importance of the fluctuations
that we have found for geodesically connected points encourages us to
pursue this research further.
\section*{ACKNOWLEDGMENTS}
It is a pleasure to thank Esteban Calzetta, Larry Ford, Jaume Garriga,
Bei-Lok Hu, Rosario Mart\'\i n and Xavier Montes for useful comments and
stimulating discussions. This work has been partially supported by the
CICYT
Research Project number AEN95-0590, and the European Project number
CI1-CT94-0004. A.R. also acknowledges support of a grant from the
Generalitat de Catalunya.
|
\section{Introduction}
Our session was given a rather general title:
"Weak Decays, CKM and CP Violation".
It is a big field and we can not do justice to
any of these subjects if we try to cover everything. For this reason,
we decided to concentrate
our discussion on penguin physics.
Year 1998 was a very good year for penguin physics -
year 1999 promises to be even better for flavor physics in general.
\begin{enumerate}
\item
We are supplied
with branching ratios on hadronic two body decays
from CLEO \cite{cleo}:
\begin{eqnarray}
{\rm Br}(B^\pm\rightarrow K^\pm\pi^0)&=&(1.5\pm 0.4\pm 0.3)\times 10^{-5},\nonumber\\
{\rm Br}(B^\pm\rightarrow K\pi^\pm)&=&(1.4\pm 0.5\pm 0.2)\times 10^{-5},\nonumber\\
{\rm Br}(B\rightarrow K^\mp\pi^\pm)&=&(1.4\pm 0.3\pm 0.1)\times 10^{-5},\nonumber\\
{\rm Br}(B\rightarrow \pi^\mp\pi^\pm)&=&\left(0.37{{+0.20}\atop{-0.17}}\right)\times 10^{-5},\nonumber\\
{\rm Br}(B\rightarrow \pi^\mp\pi^0)&=&\left(0.59{{+0.32}\atop{-0.27}}\right)\times 10^{-5}.
\mlab{CLEO}
\end{eqnarray}
\item
New results on
$\frac{\epsilon'}{\epsilon}$ was promised, and in fact, just after the meeting
the result from E832 \cite{E832} was announced. The result is considerably larger than that of previous Fermilab result\cite{E731}:
$${\rm Re} \frac{\epsilon'}{\epsilon} = \left\{
\begin{array}{ll}
(23\pm 6.5)\times 10^{-4}&{\rm NA31}
\\
(28.0 \pm 0.30_{stat.}\pm 0.26_{syst.}\pm 0.10_{{\rm MC} stat.} )
\times 10^{-4}&{\rm E832}
\end{array}
\right.$$
This result is now in agreement with that from CERN \cite{NA31}
and establishes non-vanishing direct {\bf CP}~violation.
\item Belle and Babar collaborations should start taking data on much
anticipated large CP violation in B decays. Along the way, there will get lots of data on B decays.
\item
A new K meson factory at Da$\Phi$ne should start taking data this year.
\end{enumerate}
Who knows, by year 2001, we may have a positive signal for New Physics.
Much of above experimental development demands better theoretical understanding of penguins.
\section{How big are penguins?}
Let us illustrate the importance of penguins by giving a hand-waving argument
based on the experimental result \mref{CLEO}. Less hand-waving argument is
presented in Gronau's contribution \cite{gro2}.
\begin{figure}[t]
\psfig{figure=africa-fig2.eps,height=3.3in}
\caption{Quark diagrams for
$B^+\rightarrow K^+\pi ^0$ and
$\pi^+\pi ^0$ decays.
(a) The tree graph contribution.
(b) Penguin contribution.\mlabf{fig1}}
\end{figure}
The amplitudes for tree and penguin contributions for
$K\pi$ decay mode are:
\begin{eqnarray}
T(K\pi)&=& \frac{G_F}{\sqrt{2}}{\bf V}_{ub}^*{\bf V}_{us}
[C_1(\mu)Q_{s1}^u(K\pi)+C_2(\mu)Q_{s2}^u(K\pi)]\nonumber\\
P(K\pi)_c&=&\frac{G_F}{\sqrt{2}}{\bf V}_{cb}^*{\bf V}_{cs}
[C_1(\mu)Q_{s1}^c(K\pi)+C_2(\mu)Q_{s2}^c(K\pi)]\nonumber\\
P(K\pi)_t&=&\frac{G_F}{\sqrt{2}}(-{\bf V}_{tb}^*{\bf V}_{ts})
\displaystyle\sum_{i=3}^{10}C_i(\mu)Q_{si}(K\pi).
\end{eqnarray}
To simplify our notation, set:
\begin{equation}
T(K\pi)= {\bf V}_{ub}^*{\bf V}_{us}{\cal T};~
P(K\pi)_c={\bf V}_{cb}^*{\bf V}_{cs}{\cal P}^c;~
P(K\pi)_t=-{\bf V}_{tb}^*{\bf V}_{ts}{\cal P}^t.
\mlab{3}
\end{equation}
If penguin diagram gave negligible contribution, the entire two body decays
occur through \mreff{fig1}(a). Then $B\rightarrow\pi\pi$ decay goes through a diagram in which the $s$ quark in \mreff{fig1}(a) is replaced by a $d$ quark.
For a rough argument, we ignore SU(3) breaking in the hadronic matrix elements.
Then, $B\rightarrow \pi\pi$ is given by
$T(\pi\pi)\sim\lambda^3{\cal T}$.
Since $T(K\pi)/T(\pi\pi)\sim\lambda$, we would expect:
\begin{equation}
{\rm Br}(B\rightarrow K\pi)/{\rm Br}(B\rightarrow \pi\pi)\sim{\cal O}(\lambda^2).
\end{equation}
Experimentally this is not so. This indicates that the $P(K\pi)$ amplitude is at least as large as the $T(\pi\pi)$.
For $B\rightarrow K\pi$, $P(K\pi)$ is ${\cal O}(\lambda ^2)$ and
$T(K\pi)$ is ${\cal O}(\lambda ^4)$. So, $P(K\pi)$ must be a major contributor
to the decay amplitude.
If $P(K\pi) \simeq T(\pi\pi)$, then
\begin{equation}
\frac{{\cal P}^c+{\cal P}^t}{{\cal T}}={\cal O}(\lambda),
\end{equation}
the penguin contribution is considerably larger than
what a naive estimate of the
loop graph would suggest:
\begin{equation}
\frac{{\cal P}^c+{\cal P}^t}{{\cal T}^u}
\sim
\frac{\alpha _S}{12\pi ^3}
{\rm log} \frac{m_t}{m_c} \sim {\cal O}(0.01).
\end{equation}
Gronau went through less hand-waving analysis and obtained
\begin{equation}
\frac{{\cal P}^c+{\cal P}^t}{{\cal T}^u}
=0.3\pm 0.1.
\mlab{gro}
\end{equation}
\begin{figure}[t]
\psfig{figure=fig3.eps}
\caption[]{Angles of the unitarity triangle are related to the phases of the
KM matrix. The right hand rule gives the positive direction of the
angle between two vectors.}
\mlabf{unitangle}
\end{figure}
Large penguin contribution is not always welcome. For example, they
play a role in so called "penguin pollution" which causes hadronic uncertainty
in determining $\phi_2$ and $\phi_3$. For notations see \mreff{unitangle} \footnote{ Here we use the notation which was introduced
when the unitarity triangle was first discussed in the context we
use today \cite{srs}.}. Problem penguins cause, however, does not compare
with richness they bring to flavor physics.
Unlike in K decays where effects of tree graphs dominate, in B physics,
quantum loop effects via penguins is often a leading
contribution. This gives us a window of opportunity to look for effects beyond the standard
model - as they are likely to contribute through loop effects.
Anticipating this possibility, we had the following discussions:
\begin{enumerate}
\item Reviews of penguins in B decays by M. Gronau.
\item New remarks on the determination of $\phi_1$ and $\phi_2$ by L. Oliver.
\item A critical look at $\phi_3$ from $B\rightarrow K\pi$ by R. Fleischer.
\item Model independent anlysis of $B\rightarrow K\pi$ decays and bounds on the weak
phase $\phi_3$ by M. Neubert.
\item Analysis of $B\rightarrow\eta K(K^*)$ and $B\rightarrow\eta' K(K^*)$ by D-D. Du.
\item Effects of SUSY particles in B and K decays by A. Masiero.
\item Chiral methods and predictions for $K\rightarrow\pi\pi$ by E. Paschos.
\end{enumerate}
\section{Taming the penguins}
How to get around the penguin pollution and extract the value of $\phi_2$
has been reviewed by Gronau. Oliver has presented an alternative approach
which may be useful. In his approach, $\phi_2$ is expressed as a function of
experimentally measurable quantities in $B\rightarrow\pi\pi$ decay, plus one other
parameter. It can be, for example $\frac{\cal P}{\cal T}$, obtained from
$B\rightarrow K\pi$, mentioned above.
The time dependent $B(\overline B)\rightarrow\pi^+\pi^-$ decay rates are given by:
\begin{eqnarray}
\Gamma\left({{\overline B}\atop B}\rightarrow\pi ^+\pi ^-\right) (t) &\sim&
\bigg[ |A|^2+ |\overline A| ^2
\mp \left(|A|^2- |\overline A| ^2 \right) \cos (\Delta M_{B_d} t)
\nonumber\\
&\pm& 2 {\Im} \left(\frac{q}{p} \overline A A^*\right)
\sin (\Delta M_{B_d} t) \bigg].
\mlab{10.30}
\end{eqnarray}
There are three experimental observables:
$$
|A|,~|\overline A|,~\mbox{and}~{\Im} \left(\frac{q}{p} \overline A A^*\right).
$$
Theoretically, we can write
\begin{eqnarray}
A&=&{\bf V}_{ud}{\bf V}_{ub}^*M^u+{\bf V}_{td}{\bf V}^*_{tb}M^t\nonumber\\
&=&{\bf V}_{ud}{\bf V}_{ub}^*M^u\left[1-
\left|\frac{{\bf V}_{td}{\bf V}^*_{tb}}{{\bf V}_{ud}{\bf V}_{ub}^*}\right|
e^{i\phi_2}
\frac{M^t}{M^u}\right].
\end{eqnarray}
Here $M^u={\cal T}^u-{\cal P}^c$, and $M^t={\cal P}^t-{\cal P}^c$. These are
related to previously introduced matrix elements except for the SU(3) breaking
corrections.
Theoretical unknowns are
$$
|M^u|,~\left|\frac{M^t}{M^u}\right|,~\arg\left(\frac{M^t}{M^u}\right),~\phi_2.
$$
Since there are 4 theoretical parameters and only 3 experimental observables,
we can not solve for $\phi_2$. We can, however solve for $\phi_2$ as a function
of, {\it e.g.~} $\left|\frac{M^t}{M^u}\right|$. We can, most likely obtain this
parameter from \mref{gro} looking at $B\rightarrow K\pi$ decays.
Further study is necessary to
see how the error from SU(3) symmetry breaking will affect the determination of
$\phi_2$. Also, there are some ambiguities coming from the sign of a square root
as well as that coming from $\phi_2\pm\pi$. For details see Oliver's talk
\cite{charl}.
\section{Getting maximum out of $B\rightarrow K\pi,~\pi\pi$ decays}
The CLEO collaboration has recently reported the observation of $B\rightarrow K\pi$
decays given in \mref{CLEO}.
It is clearly important to understand what we can learn from these results.
Contributions from Fleischer, Gronau, and Neubert on this subject are rather
technical. Nevertheless, it is an important technicality, as it must be understood when information is extracted from data. So, rather than summarizing what they have presented, I have presented necessary formalism which I hope is useful
in following their work.
Feynman graphs shown in \mreff{fig1} illustrates the class of operators
which are generated by QCD and electroweak radiative corrections.
The weak Hamiltonian which causes these decays can be written as
\begin{eqnarray}
{\cal H}
&=&\frac{G_F}{\sqrt{2}}\biggl\{\xi_u[C_1(\mu)O_1^u+C_2(\mu)O_2^u]
+\xi_c[C_1(\mu)O_1^c+C_2(\mu)O_2^c]\nonumber\\
&-&\xi_t\sum_{i=3}^{10}C_i(\mu)O_i\biggr\}
+\hbox{h.c.,}
\mlab{12}
\end{eqnarray}
where $\xi_q={\bf V}^*_{qb}{\bf V}_{qs}$, $O_1^q=(\bar bs)_{V-A}(\bar qq)_{V-A}$ and
$O_2^q=(\bar bq)_{V-A}(\bar qs)_{V-A}$,
$O_9=\frac{3}{2}(\bar bs)_{V-A}\sum_qe_q(\bar qq)_{V-A}$,
$O_{10}=\frac{3}{2}(\bar bq)_{V-A}\sum_q e_q(\bar qs)_{V-A}$.
Let us try to understand the isospin structure of these operators.
The up tree graph \mreff{fig1}(a) contains $u$ and $\bar u$ quarks
and they generate both $\Delta I=0$, and $\Delta I=1$ terms in the
effective Hamiltonian.
\mreff{fig1}(b), the charm tree graph, contains all isosinglet quarks and thus they generate
$\Delta I=0$ operator. \mreff{fig1}(c), the penguin, gives contribution which is proportional to $\sum_{q=u,d,s,c}\bar q\lambda^aq$ and it gives only $\Delta I=0$ operator.
Finally, \mreff{fig1}(d), the electroweak penguin, gives both $\Delta I=0$, and $\Delta I=1$ operators.
Now we consider the isospin properties of the up tree, the operator which is generated by \mreff{fig1}(a). Because it contains both
$\Delta I=0$, and $\Delta I=1$ components, we want to separate the operator into two parts:
\begin{eqnarray}
2[C_1(\mu)O_1^u+C_2(\mu)O_2^u]&=&C_+(\mu)[O_+^u-O_+^d]+C_-(\mu)[O_-^u-O_-^d]\nonumber\\
&+&C_+(\mu)[O_+^u+O_+^d]+C_-(\mu)[O_-^u+O_-^d]
\end{eqnarray}
where $C_\pm=C_2\pm C_1$ and $O_\pm=\frac{1}{2}(O_2\pm O_1)$.
Then the first two terms on the right hand side cause $\Delta I=1$ transition and the last two terms cause $\Delta I=0$ transition.
Next we show that the electroweak penguins, can be expressed interms of existing operators \cite{nr,flei}.
Note that the standard model predicts that $C_{7,8}(m_b)$ are very small and
they are negligible compared to $C_{9,10}(m_b)$.
The operators with dominant coefficients $O_9$ and $O_{10}$, for $q=u,~d$,
can be written as linear combinations of
$O_1^{u,d}$ and $O_2^{u,d}$ respectively:
\begin{eqnarray}
{\cal H}_0
&=&\frac{G_F}{\sqrt{2}}\biggl\{
\xi_c[C_+(\mu)O_+^c+C_-(\mu)O_-^c]
+
[\xi_uC_+(\mu)-\frac{1}{2}\xi_tC_+^{EW}(\mu)]\hat O_+\nonumber\\
&+&[\xi_uC_-(\mu)-\frac{1}{2}\xi_tC_-^{EW}(\mu)]\hat O_-
-\xi_t\sum_{i=3}^6C_i(\mu)O_i
\biggr\}\nonumber\\
{\cal H}_1
&=&\frac{G_F}{\sqrt{2}}\biggl\{
[\xi_uC_+(\mu)-\frac{3}{2}\xi_tC_+^{EW}(\mu)]\bar O_+
+[\xi_uC_-(\mu)-\frac{3}{2}\xi_tC_-^{EW}(\mu)]\bar O_-
\biggr\}\nonumber\\
\mlab{14.34}
\end{eqnarray}
where ${\cal H}_I$ denotes the Hamiltonian which transforms as isospin I.
The operators above are defined as:
$$
\begin{array}{lll}
\bar O_\pm=\frac{1}{2}(O^u_\pm-O^d_\pm),&~~~~~&
\hat O_\pm=\frac{1}{2}(O^u_\pm+O^d_\pm),\\
O^c_\pm=\frac{1}{2}(O_2^c\pm O_1^c),&~~~~~&
C_\pm^{EW}=C_{10}\pm C_9.
\end{array}
$$
In studying $B\rightarrow K\pi~\mbox{and~}\pi\pi$ decays, it is important to classify
final states in terms of strong interaction eigenstates, {\it i.e.}~ isospin states.
This will allow us to take in to account of all rescattering effects which
have been discussed extensively in the literature.
$$
\begin{array}{rll}
A(B^+\rightarrow\pi^+K^0)&=B_\frac{1}{2}+A_\frac{1}{2}+A_{\frac{3}{2}}&=P+A-\frac{1}{3}P_{EW}^c\\
-\sqrt{2}A(B^+\rightarrow\pi^0K^+)&=B_\frac{1}{2}+A_\frac{1}{2}-2A_{\frac{3}{2}}&
=P+T+C+A+\frac{2}{3}P_{EW}^c+P_{EW}\\
-A(B^0\rightarrow\pi^-K^+)&=B_\frac{1}{2}-A_\frac{1}{2}-A_{\frac{3}{2}}&=P+T+\frac{1}{3}P_{EW}^c\\
\sqrt{2}A(B^0\rightarrow\pi^0K^0)&=B_\frac{1}{2}-A_\frac{1}{2}+2A_{\frac{3}{2}}&=P-C
-\frac{2}{3}P_{EW}^c-P_{EW}
\end{array}
$$
where $A_I={\langle} (K\pi)_I|{\cal H}_1|B{\rangle}$ and
$B_\frac{1}{2}={\langle} (K\pi)_\frac{1}{2}|{\cal H}_0|B{\rangle}$.
We have also given the decay amplitudes in terms of amplitudes classified by
Feynman graph structure: tree graph (T), color suppressed tree graph (C), annihilation graph (A), penguin graph (P), electroweak penguin graph ($P_{EW}$), and
color suppressed electroweak penguin graph ($P_{EW}^c$).
These decays together with their charge conjugate version constitue 8 physical
observables. Unlike $K\rightarrow\pi\pi$ decays Watson's theorem cannot be applied, and
we cannot say much about final state interaction phases for these amplitudes.
We thus write \cite{nr}:
\begin{equation}
A_i=a_ie^{i\alpha_i}+b_ie^{i(\beta_i+\phi_3)}~~~~~~~~i=1,2,3.
\mlab{11}
\end{equation}
Here we separate the contributions which are proportional to
$\xi_u\equiv |\xi_u|e^{i\phi_3}$ from
those proportional to $\xi_c$ and $\xi_t$. $\alpha_i$ and $\beta_i$ are
final state strong interaction phases.
It is trivial to write $a_i$ and $b_i$ in terms of matrix elements of the effective Hamiltonian \mref{14.34}. Note that there are 12 independent parameters in
\mref{11}, and only 8 decay rates for $B\rightarrow K\pi$.
In terms of matrix elements of the Hamiltonian, we have
\begin{equation}
B_{\frac{1}{2}}\equiv P;~~~
A_\frac{3}{2}=\bar C_+\bar Q_+^\frac{3}{2}+\bar C_-\bar Q_-^\frac{3}{2};~~~
A_\frac{1}{2}=\bar C_+\bar Q_+^\frac{1}{2}+\bar C_-\bar Q_-^\frac{1}{2},
\mlab{c16}
\end{equation}
where
\begin{equation}
\bar C_\pm=\xi_u C_\pm-\frac{3}{2}\xi_tC_\pm^{EW}=|\xi_u| C_\pm(e^{i\phi_3}-\delta_\pm),
\end{equation}
and $\bar Q^I_\pm$ is an
appropriate matrix element $\mat{(K\pi)_I}{\bar O_\pm}{B}$.
We also record
\begin{equation}
P=\hat C_+\hat Q_+^\frac{1}{2}+\hat C_-\hat Q_-^\frac{1}{2}+\xi_c[C_+Q^c_++C_-Q_-^c]+\xi_t\sum_{i=3}^6C_iQ_i
\mlab{gro}
\end{equation}
where
$$
\hat C_\pm=|\xi_u| C_\pm(e^{i\phi_3}-\frac{1}{3}\delta_\pm),~~
\hat Q_\pm^\frac{1}{2}=\mat{(K\pi)_\frac{1}{2}}{\hat O_\pm}{B},~~\mbox{and}~~
Q^c_\pm=\mat{(K\pi)_\frac{1}{2}}{\hat O_\pm^c}{B}.
$$
So far, we have been quite general. Now, we shall go on to discuss the the hadronic matrix elements.
What do we know about these matrix elements? Over the years, we have learned quite a bit. In particular, we have learned that classifying operators in terms of their topology, A, C, T, P, $etc.$, gives us fairly accurate intuition in guessing the size of the matrix elements. For example, we guess that A, the annihilation graph, should be
quite small compared to T because it is suppressed by the small probability that the spectator quark and b quark come within the
range so that they can annihilate. Similarly, C is suppressed compared to T because of the color factor. These statements imply definite relationships between hadronic matrix elements which appear in \mref{c16}. These
relations should be checked experimentally. But, for the time being, we shall
proceed and ask if these conventional wisdom would allow us to determine
$\phi_3$, the KM phase.
The first simplification is
$A(B^+\rightarrow\pi^+K^0)=B_\frac{1}{2}+A_\frac{1}{2}+A_{\frac{3}{2}}\approx |P|e^{i\phi_P}$ if A and $P_{EW}$
is negligible compared to P.
The second simplification is that $\bar Q_-$ transforms like a $\Delta I=0$ operator in the limit of U spin symmetry. So, $\bar Q_-^\frac{3}{2}$ vanishes in the SU(3) limit and is
proportional to SU(3) breaking interaction. For our purpose, we neglect it. Then
$-3A_\frac{3}{2}=|P|\epsilon_{3/2}e^{i\phi_{3/2}}(e^{i\phi_3}-\delta_+)$,
where $|P|\epsilon_{3/2}=-3|\xi_u| C_+\bar Q_+^\frac{3}{2}$.
These considerations make analysis of $B^+\rightarrow K^+\pi^0$ simple:
\begin{equation}
-\sqrt{2}A(B^+\rightarrow K^+\pi^0)=A(B^+\rightarrow K^0\pi^+)-3A_\frac{3}{2}
\end{equation}
Neubert considers \cite{nr2}
\begin{equation}
R_*^{-1}=\frac{2[{\rm Br}(B^+\rightarrow K^+\pi^0)+{\rm Br}(B^-\rightarrow K^-\pi^0)]}
{{\rm Br}(B^+\rightarrow K^0\pi^+)+{\rm Br}(B^-\rightarrow K^0\pi^-)}=1-2\epsilon_{3/2}\cos\Delta\phi(\cos\phi_3-\delta_+)
\mlab{Rstar}
\end{equation}
where $\Delta\phi=\phi_\frac{3}{2}-\phi_P$.
Fleischer considers \cite{flei2}
\begin{equation}
R=\frac{[{\rm Br}(B^0\rightarrow K^+\pi^-)+{\rm Br}(\overline B^0\rightarrow K^-\pi^+)]}
{{\rm Br}(B^+\rightarrow K^0\pi^+)+{\rm Br}(B^-\rightarrow \overline K^0\pi^-)}.
\end{equation}
The decay $B^0\rightarrow K^+\pi^-$ is bit more complicated because we have to
confront the contribution from $A_\frac{1}{2}$. It involves two complex amplitudes.
He suplements the complexity by also considering
\begin{equation}
A_0=\frac{{\rm Br}(B^0\rightarrow K^+\pi^-)-{\rm Br}(\overline B^0\rightarrow K^-\pi^+)]}
{{\rm Br}(B^+\rightarrow K^0\pi^+)+{\rm Br}(B^-\rightarrow \overline K^0\pi^-)}
\end{equation}
Detailed numerical analysis indicates that both of these methods are useful
for determining $\phi_3$.
In this discussion, I had to simplify the
problem in order to present the essence. I refer the reader to the original
contribution for complete analysis.
It is clear that their contributions lead to much progress but much more
work is necessary along this direction.
For example, only subset of $B\rightarrow K\pi$ has been
considered. There are 8 of them altogether!
\section{SUSY in $B$ and $K$ decays}
Predictions of the minimal supersymmetric theories (MSSM)
is essentially same as those of the SM. If nature has
chosen MSSM, we will not learn anything new from experiments
on $B$ and $K$ decays. We should not be too discouraged by this
though, as it is likely that she has chosen a theory which
is more elegant than the MSSM. But, as long as we can not
specify which theory nature has chosen, it is not an easy task to analyze its
predictions. There are as many as 124 parameters in a non-
minimal SUSY, and perhaps even more.
Because $B$ and $K$ decays give stringent restrictions
on flavor changing neutral current strengths, we shall
focus general predictions of FCNC processes of a non-minimal
SUSY - mostly penguin effects.
In SUSY, there is a bosonic partner for each helicity of
a quark. Here we begin by examining a $6\times 6$ squark mass matrix
of the MSSM.
\begin{equation}
{\bf \tilde M}^2_D=\pmatrix{
{\bf \tilde M}_{DLL}^{\rm tree}{\bf \tilde M}_{DLL}^{\rm tree\dagger}
+c_1{\cal M}_U{\cal M}_U^\dagger
&Am_{3/2}{\cal M}_D(1+\frac{c_2}{M_W^2}
{\cal M}^\dagger_U{\cal M}_U)\cr
A^*m_{3/2}(1+\frac{c_2}{M_W^2}
{\cal M}^\dagger_U{\cal M}_U){\cal M}^\dagger_D
&{\bf \tilde M}_{DRR}^{\rm tree\dagger}{\bf \tilde M}_{DRR}^{\rm tree}}
\mlab{massmw}
\end{equation}
where
\begin{eqnarray}
{\bf \hat M}^2_{DLL}&=&\left( m^2_{3/2}+(v_1^2-v_2^2)
\left(\frac{{g'}^2}{12}- \frac{g^2}{4}\right)\right){\bf 1}+
({\cal M}^{\rm diag}_D)^2\nonumber\\
{\bf \hat M}^2_{DRR}&=&\left(m^2_{3/2}+
(v_1^2-v_2^2)\frac{{g'}^2}{6}\right){\bf 1} +({\cal M}^{\rm diag}_D)^2\nonumber\\
{\bf \hat M}^2_{DLR}&=&\left(|A|m_{3/2}+\mu^*\frac{v_1}{v_2}\right)
{\cal M}^{\rm diag}_D.
\mlab{massplanck}
\end{eqnarray}
${\bf 1}$ is a $2\times 2$ unit matrix, ${\cal M}_{D,(U)}$ is a mass matrix of a $D(U)$ quark, and
${\cal M}^{\rm diag}_{D,(U)}$ is a corresponding diagonalized matrix.
Note that there new FCNC effects from additional flavor mixing among squarks.
Others are standard MSSM parameters. To go from MSSM to a more general theory, let us identify
\begin{eqnarray}
\Delta_{LL}^2&=&c_1{\cal M}_U{\cal M}_U^\dagger\nonumber\\
\Delta_{LR}^2&=&Am_{3/2}{\cal M}_D
\left( 1+\frac{c_2}{M_W^2}
{\cal M}^\dagger_U{\cal M}_U \right) \nonumber\\
\Delta_{RL}^2&=&A^*m_{3/2}\left( 1+\frac{c_2}{M_W^2}
{\cal M}^\dagger_U{\cal M}_U\right) {\cal M}^\dagger_D.
\mlab{DELTAAB}
\end{eqnarray}
and generalize $\delta_{LL}=\frac{\Delta_{LL}}{\tilde m}$,
$\delta_{LR}=\frac{\Delta_{LR}}{\tilde m}$,
$\delta_{RL}=\frac{\Delta_{RL}}{\tilde m}$
, and $\delta_{RR}=\frac{\Delta_{RR}}{\tilde m}$, where
$\tilde m$ is an average squark mass,
as new arbitrary dimensionless parameters.
We then study experimental constraints on $\delta$ ignoring all
theoretical prejudice.
This analysis has been discussed in detail by Masiero.
Bounds on $\delta$ has been obtained from experimental
information on various FCNC processes. For
an average squark mass and gluino mass of $500$GeV,
the bounds on $\delta$ ranges from $10^{-1}$ to
$10^{-3}$. It sould be noted that the neutron edm
gives a bound of $\Im \left(\delta_{LR}\right)_{11}\sim 10^{-6}$
and it is natural to assume that other components of $\delta$
are of the same order of magnitude. If this is the case, it may be
difficult of SUSY effects to show up in B and K decays.
\section{B decays to $\eta K(K^*)$, $\eta'K(K^*)$, and
$\eta'X_s$}
CLEO has observed \cite{cleoeta}
\begin{eqnarray}
{\rm Br}(B\rightarrow \eta'K^+)&=&\left(6.5{{+1.5}\atop{-1.4}}\pm 0.9\right)\times 10^{-5},\nonumber\\
{\rm Br}(B\rightarrow \eta'K^0)&=&\left(4.7{{+2.7}\atop{-2.0}}\pm 0.9\right)\times 10^{-5}.\nonumber\\
\end{eqnarray}
These branching ratios are surprisingly large. Du has presented
a review of work in progress to undersand these large
branching ratios. It is likely that these branching ratio is large because
of gluonic content of $\eta'$.
Among various suggestions, a particularly
interesting one is that of Soni and Atwood. They attempt
to compute $\eta'\rightarrow$ glue glue by considering triangle
anomaly \cite{as}. They then estimate $b\rightarrow\eta'$ gluon.
We should note, however, that major contribution to this decay comes from
off shell gluon. Thus the validity of the anomaly calculation is questionable
at best. A universal characteristic of all the work presented by Du is that
each author picks up their favorite diagram and estimates its contribution.
Nature does not work this way. They have to consider all possible diagrams
and sum them up. Clearly global analysis is urgently needed.
Also, there are large amount of data on the gluonic content of $\eta'$.
Such global analysis must be consistent with the previously known properties
of $\eta'$.
\section{A new calculation for direct {\bf CP}~violation:
$\frac{\epsilon'}{\epsilon}$}
Ever since the discovery of {\bf CP}~violation, experimentalists have
been looking for an evidence of direct {\bf CP}~violation.
Now that the result of NA31 has been confirmed by E832,
the direct {\bf CP}~ violation has been experimentally established.
The challenge for theorists is to extract physics from
the new value of $\frac{\epsilon'}{\epsilon}$. This
is not an easy task. Before we compute {\bf CP}~violating
amplitudes for $K\rightarrow\pi\pi$ decay, we have to demonstrate
that we understand {\bf CP}~conserving decay amplitudes.
This means that we need to understand the $\Delta I=\frac{1}{2}$
rule.
At the moment there is no clear understanding of this rule.
So, one choice is to obtain hadronic matrix elements
from data \cite{buras}. If the
$\Delta I=\frac{1}{2}$ rule is due to some new physics, this procedure
will miss the new physics. Clearly, this is not satisfactory.
We want to compute everything from basic principles.
The approach taken by Paschos is an attempt along this direction.
Let us start from the defining equation:
\begin{equation}
\epsilon ^{\prime} =e^{i(\delta_2-\delta_0)}
\frac{1}{2\sqrt{2}}\omega
\left( \frac{\overline A_0}{A_0} - \frac{\overline A_2}{A_2}\right) .
\mlab{6.107}
\end{equation}
where $A_I=\mat{(\pi\pi)_I}{{\cal H}}{K}$ and $\omega=|A_2/A_0|\sim .05$
and $\delta_I$ is the $\pi\pi$ phase shift for isospin $I$ channel.
In terms of operators in the effective Hamiltonian ${\cal H}$,
\begin{equation}
\frac{\epsilon'}{\epsilon}
=-ie^{i(\delta_2-\delta_0)}\times 10^{-4}
\left(\frac{\Im\tau}{10^{-4}}\right)
r\left(\sum_{i}y_i{\langle} Q_i{\rangle} _0-\frac{1}{\omega}\sum_{i}
y_i{\langle} Q_i{\rangle} _2\right)
\mlab{epe8}
\end{equation}
where $r=\frac{G_F\omega}{|2\epsilon|\Re A_0}=336~{\rm GeV}^{-3}$;
$y_i$ is the imaginary part of Wilson coefficients;
${\langle} Q_i{\rangle} _I$ is a hadronic matrix element
$\mat{(\pi\pi)_I }{0_i}{K}$;
$\tau=-\frac{{\bf V}^*_{ts}{\bf V}_{td}}{{\bf V}^*_{us}{\bf V}_{ud}}$.
In this workshop, Paschos described computation of hadronic
matrix elements ${\langle} Q_i{\rangle} _I$ based on
the $\frac{1}{N_C}$ expansion.
They have obtained
\begin{equation}
5\times 10^{-4}\leq \frac{\epsilon'}{\epsilon}\leq 22\times 10^{-4}
\end{equation}
for $m_s=(150\sim 175)$MeV. Their prediction increases if smaller
values of $m_s$ is taken.
This is an on going study and much more work remains:
(I)The Wilson coefficients can be computed reliably only
at some large energy scale $\Lambda_c$. But hadronic matrix elements
can be computed only at low energy scale. So, there has to be some compromise.
They ave chosen $\Lambda_c\sim 800\mbox{MeV}$. Wilson coefficient
functions have large scale dependence in this region. When the coefficient
functions and hadronic matrix elements are combined, the result for
$\frac{\epsilon'}{\epsilon}$ should not have the scale dependence. This
has to be studied carefully.
(II) It is necessary to
demonstrate that $K\rightarrow\pi\pi$ decay can be understood in this framework.
Indeed, their result for $K^+\rightarrow\pi^+\pi^0$ decay, the $\Delta I=\frac{3}{2}$
channel, is consistent with experiment. But it is necessary to understand
the $\Delta I=\frac{1}{2}$ amplitude. Personally, I am skeptical toward
a claim that the $\Delta I=\frac{1}{2}$ rule can be understood within the frame
work of $\frac{1}{N_C}$ expansion.
\section{Summary}
We have tried to have extended discussions on penguins.
We tried to understand how we might extract $\phi_2$ and $\phi_3$ from data.
We don't think there is one best method to extract these angles.
It is an experiment driven field, and time will tell. We tried to understand
$\frac{\epsilon'}{\epsilon}$ from basic principles of the SM. There is much more
work to be done along this direction. We tried to understand $B\rightarrow\eta'+X$ decays. This also requires more work. Seeing effects of SUSY in $B$ and $K$ decays
is an exciting possibility. We have to keep on searching.
\vskip 1cm
\centerline{Acknowledgements}
This work has been supported in part by Grant-in-Aid for Special Project
Research
(Physics of {\bf CP}~violation). Comments from J. L. Rosner was helpful in finalizing the manusacript. I wish to express my gratitude to the organizers,
C. A. Dominguez and R. D. Viollier for their effort in organizing the
workshop, and for their hospitality.
\section*{References}
|
\section{Introduction}
The reduction of membrane theory can lead to a simple model,
describing an isentropic, irrotational fluid \cite{BH}.
A similar system can be obtained, e. g.,
by dimensional reduction of relativistic field theory
\cite{JEV},
and also in the hydrodynamical formulation
of the (non--linear) Schr\"odinger equation \cite{BJ}.
The model, also used in gas dynamics,
was further discussed by
Bazeia, Jackiw, and Polychronakos
\cite{BJ}, \cite{JP1}, \cite{BAZ}, \cite{JP2};
note also \cite{NUTKU}.
Let us consider the action
\begin{equation}
{\cal S}=\int\!dxdt\underbrace{\Big[- R{\partial}_{t}\Theta
-\frac{1}{2} R(\partial_{x}\Theta)^2}_{{\cal L}_{0}}
-V( R)\Big],
\label{BJ}
\end{equation}
where $ R(x,t)\geq0$ and $\Theta(x,t)$ are real fields
and $V(R)$ is some potential. (Our Lagrange density
differs from that of Bazeia and Jackiw in \cite{BJ}
in a surface term;
the two expressions are hence equivalent.
Albeit similar results hold in any dimension, we shall
restrict ourselves, for simplicity, to $(1+1)$ dimensionsal
space--time, parametrized by position and time, $x$ and $t$.).
The associated Euler-Lagrange equations read
\begin{equation}
\partial_{t} R+\partial_{x}\big( R\partial_{x}\Theta\big)=0,
\qquad
\partial_t\Theta+\frac{1}{2}(\partial_x\Theta)^2=-\frac{dV}{dR}.
\label{eqmotion}
\end{equation}
In what follows, we shall (except in Section 7), restrict
ourselves
to potentials of the form
$
V=cR^\omega,
$
where $c$ and $\omega$ are real constants.
In the membrane case the effective potential is in particular
\begin{equation}
V(R)=\frac{c}{R},
\qquad
c={\rm const}.
\label{membpot}
\end{equation}
The Lagrangian (\ref{BJ}) is first-order in the time derivative;
it admits therefore an (extended)
Galilean symmetry, with conserved quantities
\begin{equation}
\begin{array}{cc}
H=\displaystyle\int dx\underbrace{\left(
\frac{1}{2} R(\partial_{x}\Theta)^2+V(R)\right)
}_{{\cal H}}\qquad\hfill
&\hbox{energy}\\
\cr
P=\displaystyle\int\! dx\underbrace{
R\,\partial_{x}\Theta}
_{{\cal P}}\hfill
&\hbox{momentum}\\
\cr
B=\displaystyle\int\! dx\,\big(x R-t{\cal P}\big)\hfill
&\hbox{boosts}\\
\cr
N=\displaystyle\int\! dx\, R\qquad\hfill
&\hbox{particle number}\\
\end{array}
\label{Galconst}
\end{equation}
Unexpectedly, the free and the membrane systems
both admit two additional conserved quantities
\cite{BH}, \cite{JEV} \cite{BJ}, namely
\begin{equation}
\begin{array}{cc}
G=\displaystyle\int dx\,
\big(x{\cal H}-\Theta{\cal P}\big)\hfill
&\hbox{``antiboost''}\qquad\hfill
\\
\cr
D=tH-\displaystyle\int dx\, R\Theta\qquad\hfill
&\hbox{time dilatation}\hfill
\hfill
\label{BJconst}
\end{array}
\end{equation}
The generators
(\ref{Galconst})--(\ref{BJconst}) span furthermore
the $(2+1)$ dimensional Poincar\'e algebra \cite{BH},
\cite{JEV}, \cite{BJ}.
The arisal of the typically {\it relativistic}
Poincar\'e symmetry for a {\it non-relativistic system}
is quite surprising.
The mystery is increased by that this
symmetry is {\it not} associated
to any finite--dimensional group action on space-time.
It belongs in fact
to a new type of ``field-dependent'' non-linear
action on space--time \cite{BJ} which, to our knowledge,
has never been met before.
Before explaining how these symmetries arise, we point out
that, in the ``free case'' $V=0$,
the entire conformal group ${\rm O}(3,2)$ is a symmetry;
it is reduced to the Poincar\'e group for
$V(R)=c/R$, and to the Schr\"odinger group
(the symmetry of the free Schr\"odinger equation \cite{JHN})
for $V(R)=cR^3$, respectively.
\goodbreak
Where do these symmetries come from~?
We answer this question
by unfolding the system into a higher--dimensional space, obtained
by promoting the ``phase'' $\Theta$ to a ``vertical'' coordinate
(we denote by $s$) on extended space, $M$.
Such a ``Kaluza--Klein--type'' framework for non--relativistic
physics was put forward by Duval et al. \cite{barg}.
In our case, their extended space
$M$ is $(2+1)$--dimensional Minkowski space,
with $x$ a spacelike and $t$ and $s$ light--cone coordinates.
Then the strange, field--dependent, non--linear action
of Bazeia and Jackiw \cite{BJ}, (Eq. (\ref{BJtransf}) below),
becomes the natural, linear action of the $(2+1)$--dimensional
Poincar\'e group on extended space.
Our starting point is the simple but crucial observation due
to Christian Duval \cite{Duval} which says that, on extended space
$M$, the ``antiboosts'' are
the counterparts of galilean boosts, when galilean time, $t$, and
the ``vertical coordinate'', $s$ are interchanged,
\begin{equation}
t\longleftrightarrow s.
\end{equation}
Many results presented in this paper come by exploiting
this intechange--symmetry.
For example,
\vskip2mm
{$\bullet$} applied to the Galilei group,
the Poincar\'e group is obtained;
{$\bullet$} applied to ``non-relativistic
conformal symmetries'' (Eq. (\ref{nrctransf}) below) yields
relativistic
conformal symmetries,
\vskip2mm\noindent
etc.
It also provides a clue for the non--conventional
implementations on fields.
\goodbreak
The action of the conformal group ${\rm O}(3,2)$ and its various
subgroups on $M$ is presented in Section 4. In Section 5
we project the natural, linear action on extended space to
a ``field-dependent action'' on ordinary space. This requires to
generalise as in Eq. (\ref{newequiv}) the
usual equivariance condition (\ref{oldequiv}) of Duval et
al. \cite{barg}.
The authors of Ref. \cite{BJ}
call the Poincar\'e symmetry ``dynamical''
since it is not associated to a natural ``geometric''
action on space--time. Our point is that these symmetries become
``geometric'' on extended space.
In Section 6 we study physics in the extended space
and show how the previous results can be recovered.
Our results show also that the
``membrane potential''(\ref{membpot}) i.e. $V(R)=c/R$ is
the only one which
can accomodate these new type of symmetries.
This is the reason why these strange
symmetries do {\it not} arise for the ordinary Schr\"odinger
equation~:
this latter corresponds in fact to a particular effective
potential, namely to
$
\overline{V}=-\frac{1}{8}
\frac{(\vec{\nabla} R)^2}{R}.
$
Usual equivariance allows us in turn to recover the
well-known Schr\"odinger symmetry.
In Ref. \cite{JP2}, the Poincar\'e symmetry of the fluid system
(\ref{BJ}) is related to that of the Nambu-Goto action of a
membrane
moving in higher dimensional space-time.
Our ``Kaluza--Klein'' framework is an
alternative way of obtaining the same conclusion.
\goodbreak
\section{Symmetries}
We first recall the construction of the conserved quantities.
Let us consider a non-relativistic theory given by the Lagrange
density ${\cal L}(\partial_{\alpha}\phi,\phi)$, where $\phi$ denotes
all fields collectively. Then Noether's theorem \cite{JM} says that
if the Lagrange density changes by a surface term
under an infinitesimal transformation
$\phi\to\phi+\delta\phi$,
\begin{equation}
\delta{\cal L}={\partial}_{\alpha}C^\alpha,
\end{equation}
then
$
J^\alpha=\frac{\delta{\cal L}}{\delta({\partial}_{\alpha}\phi)}
\delta\phi-C^\alpha
$
is a conserved current, ${\partial}_{\alpha}J^\alpha=0$, so that
\begin{equation}
\displaystyle{\int{\!dx\,
{\Big(\frac{\delta{\cal L}}{\delta({\partial}_{t}\phi)}\delta\phi-C^t\Big)}}}
\label{consquant}
\end{equation}
is conserved.
For example, the usual Galilean
transformations of non--relati\-vis\-tic space-time,
$\left(\begin{array}{c}
x
\\
t
\end{array}\right)
\to
\left(\begin{array}{c}
{x}^\star
\\
{t}^\star
\end{array}\right)
$,
$\left(\begin{array}{c}
R(x,t)
\\
\Theta(x,t)
\end{array}\right)
\to
\left(\begin{array}{c}
{R}^\star(x,t)
\\
{\Theta}^\star(x,t)
\end{array}\right)
$,
where
\begin{equation}
\begin{array}{ccc}
\begin{array}{c}
{x}^\star=x,\hfill
\\
{t}^\star=t+\tau,\hfill
\\
\end{array}\hfill
&
\begin{array}{c}
R^\star(x,t)=R(x,t+\tau),\hfill
\\
\Theta^\star(x,t)=\Theta(x,t+\tau);\hfill\\
\end{array}\hfill
&\hbox{time translation}
\\
\cr
\begin{array}{c}
x^\star=x-\gamma,\hfill
\\
{t}^\star=t,\hfill
\\
\end{array}\hfill
&\begin{array}{c}
R^\star(x,t)= R(x-\gamma,t),\hfill
\\
\Theta^\star(x,t)=\Theta(x-\gamma,t);\hfill\\
\end{array}\hfill
&\hbox{
translation}\\
\cr
\begin{array}{c}
{x}^\star=x+\beta t,
\hfill
\\
{t}^\star=t,\hfill
\\
\\
\end{array}\hfill
&\begin{array}{cc}
R^\star(x,t)=& R(x+\beta t,t),\hfill
\\
\Theta^\star(x,t)=&
\Theta(x+\beta t,t)\hfill
\\
&-\beta x-\frac{1}{2}\beta^2t;
\hfill
\\
\end{array}\hfill
&\hbox{boost}\\
\\
\begin{array}{c}
{x}^\star=x
\hfill
\\
{t}^\star=t,
\\
\end{array}\hfill
&\begin{array}{c}
R^\star(x,t)= R(x,t),\hfill
\\
\Theta^\star(x,t)=\Theta(x,t)-\eta.
\end{array}\hfill
&\hbox{phase shift}\\
\end{array}
\label{Galtr}
\end{equation}
change the Lagrange density (\ref{BJ}) by a surface term,
and Noether's theorem yields the conserved quantities
(\ref{Galconst}).
The new conserved quantities (\ref{BJconst}) belong in turn
to the following
strange, non--linear action on space-time \cite{BJ}
\begin{equation}
\begin{array}{cc}
\begin{array}{c}
{x}^\star=
x+\alpha\Theta({x}^\star,{t}^\star)
\hfill
\\
{t}^\star=t+\frac{1}{2}\alpha
\big(x+{x}^\star)
\\
\end{array}\qquad
\hfill
&\hbox{``antiboost''}
\\
\cr
\begin{array}{c}
{x}^\star =x
\hfill
\\
t^\star =e^{\delta}t\hfill
\\
\end{array}
\hfill
&\hbox{time dilatation}
\\
\end{array}
\label{BJtransf}
\end{equation}
``Antiboosts'' are particularly interesting~:
$x^\star $ and
${t}^\star $ are only defined implicitely,
and the action is ``field--\-depen\-dent''
in that its very definition involves $\Theta$.
When implemented on the fields non--conventionally,
these transformations act as symmetries.
In detail, let us set
\begin{equation}
\begin{array}{cc}
\begin{array}{c}
R^\star (x,t)=
\displaystyle{\frac{ R({x}^\star, {t}^\star)}{J^\star}},\hfill
\\
\Theta^\star(x,t)=\Theta({x}^\star,{t}^\star);
\qquad\hfill
\\
\end{array}\hfill
&\hbox{``antiboost''}\hfill
\\
\cr
\begin{array}{c}
R^\star(x,t)=
e^{-\delta} R({x}^\star,{t}^\star),
\qquad\hfill
\\
\Theta^\star(x,t)=
e^{\delta}\Theta({x}^\star,{t}^\star);\hfill
\\
\end{array}\hfill
&\hbox{time dilatation}\hfill
\label{BJimp}
\end{array}
\end{equation}
where
$J^\star=
\big(1-\alpha\partial_{{x}^\star}\Theta({x}^\star,{t}^\star)
-\frac{1}{2}
\alpha^2\partial_{{t}^\star}\Theta({x}^\star,
{t}^\star)\big)^{-1}$
is the Jacobian of the space-time transformation
(\ref{BJtransf}). Then
the Lagrangian (\ref{BJ}) changes by a surface term and
the conserved quantities (\ref{BJconst}) are
recovered by Noether's theorem.
So far we merely reviewed the results from Ref. \cite{BJ}.
Now we point
out that, in the free case $V=0$, the system
described by the Lagrangian ${\cal L}_{0}$ in (\ref{BJ})
has even more symmetries.
Let us first remember that
the ``non-relativistic conformal transformations''
\begin{equation}
\begin{array}{cc}
\begin{array}{c}
{x}^\star =e^{\lambda/2}x,\hfill
\\
{t}^\star =e^{\lambda}t;\hfill
\\
\end{array}
\hfill
&\hbox{non-relat. dilatations}\hfill
\\
\cr
\begin{array}{c}
{x}^\star =\displaystyle\frac{x}{1-\kappa t},
\hfill
\\
{t}^\star =\displaystyle\frac{t}{1-\kappa t};\qquad
\hfill
\end{array}
&\hbox{expansions}\hfill
\\
\end{array}
\label{nrctransf}
\end{equation}
are symmetries for
the free Schr\"odinger equation \cite{JHN}.
The transformations in (\ref{nrctransf}) span
with the time translation $t^\star=t+\epsilon$
an ${\rm SL}(2,{\rm R\hspace*{-2ex} I\hspace*{1.2ex}})$ group;
added to the Galilei transformations (\ref{Galtr}), the
Schr\"odinger group is obtained.
Implementing (\ref{nrctransf}) on $R$ and $\Theta$ as
\begin{equation}
\begin{array}{cc}
\begin{array}{c}
R^\star(x,t)=
\displaystyle{e^{\lambda/2}} R({x}^\star,{t}^\star),
\hfill
\\
\Theta^\star(x,t)=
\Theta({x}^\star,{t}^\star);\hfill
\\
\end{array}\hfill
&\hbox{non-relat. dilatation}\hfill
\\
\\
\begin{array}{c}
R^\star(x,t)=
\displaystyle{\frac{1}{1-\kappa t}}\, R({x}^\star,{t}^\star),
\hfill
\\
\Theta^\star(x,t)=
\Theta({x}^\star,{t}^\star)-
\displaystyle\frac{\kappa x^2}{2(1-\kappa t)};\hfill
\\
\end{array}\hfill
&\hbox{expansion}\hfill
\end{array}
\label{nrctrimp}
\end{equation}
the free action is left invariant.
Thus, the transformations in (\ref{nrctransf})
act as symmetries also in our case.
The associated conserved quantities read
\begin{equation}
\begin{array}{cc}
\Delta=\displaystyle\int dx\,
\big(t{\cal H}-\frac{1}{2}x{\cal P}\big),\hfill
&\hbox{non-relativistic dilatation}\hfill
\\
\cr
K=-t^2H+2t\Delta
+\frac{1}{2}\displaystyle\int dx\,
x^2 R;\qquad\hfill
&\hbox{expansion}\hfill
\hfill
\label{newnrconst}
\end{array}
\end{equation}
Remarkably, the ``relativistic''
dynamical Poincar\'e symmetry can also
be conformally extended. Using the equations of motion
(\ref{eqmotion}), a lengthy but straightforward calculation shows
that, for $V=0$,
\begin{equation}
\begin{array}{c}
{\rm C}_{1}=\displaystyle\int dx\,\Big(
\frac{x^2}{2}{\cal H}-x\Theta{\cal P}+\Theta^2 R
\Big),\hfill
\\
{\rm C}_{2}=\displaystyle\int dx\Big(
xt{\cal H}-\big(\frac{x^2}{2}+t\Theta\big){\cal P}
+x\Theta R\Big)\;\quad
\hfill
\label{newrelconst}
\end{array}
\end{equation}
are also conserved, $\frac{d{\rm C}_{i}}{dt}=0$.
A shorter proof can be obtained by calculating
the energy--momentum tensor for (\ref{BJ}),
\begin{equation}
\begin{array}{c}
T_{tt}=
\displaystyle{\frac{ R}{2}}\big(\partial_{x}\Theta\big)^2
+cR^\omega,
\hfill
\ccr
T_{xt}=- R\partial_{x}\Theta\,\partial_{t}\Theta,\hfill
\ccr
T_{tx}= R\partial_{x}\Theta,\hfill
\ccr
T_{xx}= R\big(\partial_{x}\Theta\big)^2+(\omega-1)c R^\omega.
\hfill
\\
\end{array}
\label{BJem}
\end{equation}
(Let us note that in the
non--relativistic context the
usual index gymnastics is meaningless, since space-time does
not carry
a metric. We agree therefore that the
energy--momentum tensor $T_{\alpha\beta}$ is a covariant
two tensor
carrying lower indices, while $T^{\alpha\beta}$ is not defined).
The tensor $T_{\alpha\beta}$ is neither symmetric
nor traceless. It is nevertheless conserved,
$
{\partial}_{\alpha}T_{\alpha\beta}=0
$
for all $\beta=t, x$. Let us rewrite the quantities
${\rm C}_{1}$ and ${\rm C}_{2}$ as the integrals of
\begin{equation}
\begin{array}{cc}
\displaystyle{\frac{x^2}{2}}T_{tt}^0-x\Theta T_{tx}^0+\Theta^2
R\qquad\hfill
&{\rm C}_{1},\hfill
\ccr
xtT_{tt}^0-\big(\displaystyle{\frac{x^2}{2}}+t\Theta\big)
T_{tx}^0
+x\Theta R\qquad\hfill
&{\rm C}_{2},\hfill
\\
\end{array}
\label{Ttransc}
\end{equation}
where $T_{\alpha\beta}^0$ denotes the free ($V=0$) energy-momentum
tensor.
Then the conservation of the quantities (\ref{newrelconst}) is
obtained by
deriving this expression w. r. t. time and using the continuity
equation ${\partial}_{\alpha}T_{\alpha\beta}^0=0.$
The Poisson brackets of our conserved quantities
(listed in Appendix A)
yield a closed, finite--dimensional algebra.
In the next Section we prove that this is in fact
the ${\rm o}(3,2)$ conformal algebra.
For the membrane potential $V=c/R$,
the conformal symmetries $\Delta$, $K$,
${\rm C}_{1}$ and ${\rm C}_{2}$
are broken, and only the Poincar\'e symmetry survives.
Changing the question, we can also ask for
what potentials do we have the same symmetries as in the
free case.
Now the Lagrangian (\ref{BJ}) is dilation
and indeed Schr\"odinger invariant only for
\begin{equation}
V=cR^3.
\label{confinvpot}
\end{equation}
This comes from the scaling properties of the Lagrange density,
and can also be seen of by looking at the
energy-momentum tensor (\ref{BJem})~:
the trace condition
\begin{equation}
T_{xx}=2T_{tt},
\label{tracecond}
\end{equation}
which is the signal for a Schr\"odinger symmetry \cite{JHN},
only holds
for $\omega=3$.
On the other hand, the potential $c R^\omega$ yields an
``antiboost--invariant''
expression only for $\omega=-1$ so that he
Poincar\'e symmetry only allows the ``membrane potential''
(\ref{membpot}),
$
V(R)={c}/{R}.
$
Therefore, the full ${\rm o}(3,2)$ conformal symmetry only arises in
the free case.
\goodbreak
\section{A ``Kaluza-Klein'' framework}
In order to explain the origin of the symmetries of the model,
let us start with Duval's unpublished observation \cite{Duval}.
Let us enlarge space-time by adding
a new, ``phase-like'' coordinate $s$ i.e.,
consider the ``extended space''
\begin{equation}
M=\left\{\left(
\begin{array}{c}
x\\
t\\
s
\end{array}
\right)
\right\}.
\end{equation}
Let us lift the space--time transformations
to $M$ by adding a transformation rule for $s$
inspired from the
rule the phase changes in Eq. (\ref{BJimp}).
Thus, let us formally replace
the field $\Theta^\star(x,t)$ by the coordinate $-s$,
\begin{equation}
\Theta^\star(x, t)\to-s.
\label{thetarule}
\end{equation}
When applied to an ``antiboost'', for example,
we get
the {\it linear} action on extended space\footnote{
Our notations are as follows.
$\mu, \nu, \dots = x, t, s$
are indices on the extended space $M$, and
$\alpha, \beta \dots = x, t$
are indices on ordinary space--time, $Q$. The
transformed coordinates are denoted by ``tilde''
($\widetilde{\{\,\cdot\,\}}$) on $M$,
while they are denoted by ``star'', ($\{\,\cdot\,\}^\star$),
on $Q$.
The fields on $M$ are denoted by lower--case letters
(e. g., $\rho, \, \theta$),
while the fields on $Q$ are
denoted by upper--case letters like $R,\, \Theta$.},
$\left(\begin{array}{c}
x\\
t\\
s
\end{array}\right)\to
\left(\begin{array}{c}
\widetilde{x}\\
\widetilde{t}\\
\widetilde{s}
\end{array}\right)$,
\begin{equation}
\begin{array}{cc}
G~:\qquad\hfill
&\begin{array}{c}
\widetilde{x}=
x-\alpha s,
\hfill
\\
\widetilde{t}=t+\alpha x-\frac{1}{2}\alpha^2s,
\hfill
\\
\widetilde{s}=s.\hfill
\\
\end{array}
\hfill
\end{array}
\label{BJBtransf}
\end{equation}
On the other hand, lifting the action of galilean boosts to our
extended space-time by applying the same rules, we get
\begin{equation}
\begin{array}{cc}
B~:\qquad\hfill
\begin{array}{c}
\widetilde{x}=x+\beta t,\hfill
\\
\widetilde{t}=t,\hfill
\\
\widetilde{s}=s-\beta x-\frac{1}{2}\beta^2t.\hfill
\\
\end{array}
\hfill
\end{array}
\label{GalBtr}
\end{equation}
The action of the mysterious ``antiboost'' becomes hence
analogous to that of galilean boost, the only difference being
that ordinary time, $t$, and the new, phase-like coordinate, $s$,
have to be interchanged,
\begin{equation}
t\longleftrightarrow s.
\label{tsinterchange}
\end{equation}
When interchanging $t$ and $s$,
the dilations of time alone in Eq. (\ref{BJtransf})
lifted to extended space by the same rule as above, remain
dilations of time alone but with the inverse parameter~:
$\delta\to-\delta$,
\begin{equation}
\begin{array}{cccc}
D~:\;\hfill
&\begin{array}{c}
\widetilde{x}=x,
\hfill
\\
\widetilde{t}=e^{\delta}t,\hfill
\\
\widetilde{s}=e^{-\delta}s
\end{array}
&\Longrightarrow
&\begin{array}{c}
\widetilde{x}=x,
\hfill
\\
\widetilde{t}=e^{-\delta}t,\hfill
\\
\widetilde{s}=e^{\delta}s.
\end{array}
\end{array}
\label{BJBtransf2}
\end{equation}
This same rule changes a time translation with parameter
$\epsilon=-\eta$
into a the phase translation,
\begin{equation}
\begin{array}{ccc}
\hbox{time translation}\qquad\hfill
&
&\hbox{phase translation}
\\
\widetilde{x}=x\hfill
&
&\widetilde{x}=x\hfill
\\
\widetilde{t}=
t+\epsilon\hfill
&\Longrightarrow\qquad\hfill
&\widetilde{t}=t\hfill
\\
\widetilde{s}=s
\qquad\hfill
&
&\widetilde{s}=s-\eta\hfill
\end{array}
\label{tvtr}
\end{equation}
\vskip-2mm
i.e.,
\begin{eqnarray*}
\vbox{\halign{#\qquad\qquad &#\qquad\qquad &#
\cr
energy\hfill
&$\Longrightarrow$\hfill
&particle number
\cr
}
}
\end{eqnarray*}
Our trick of adding an extra coordinate $s$ allowed us so far
to reconstruct the Poincar\'e group from the extended
Galilei group
by the ``interchange rule'' (\ref{tsinterchange}).
The conformal extensions can be similarly investigated.
Non--relativistic dilations act as
\begin{equation}
\begin{array}{ccc}
\Delta:
\left(
\begin{array}{c}
x\\
t\\
s\\
\end{array}\right)
&\to
\hfill
&\left(\begin{array}{c}
e^{\lambda/2}x,\hfill
\\
e^{\lambda}t;\hfill
\\
s
\end{array}\right).\hfill
\end{array}
\end{equation}
Let us observe that relativistic dilations, i. e.,
uniform dilations of all coordinates,
$
d:\left(
\begin{array}{c}
x
\\
t
\\
s
\end{array}\right)
\to
\left(\begin{array}{c}
e^{\delta}x
\\
e^{\delta} t\\
e^{\delta}s
\\
\end{array}\right)
$
also belong to our algebra, since they correspond to a
non-relativistic dilation ($\Delta$) with parameter
$2\delta$,
followed by
a dilation of time alone ($D$) with parameter $-\delta$,
$
d=D_{-\delta}
{\raise 0.5pt \hbox{$\scriptstyle\circ$}}\Delta_{2\delta}.
$
Then the $t\leftrightarrow s$ counterpart of a
non--relativistic dilation is
a uniform dilation followed by a dilation of time alone,
\begin{equation}
\Delta_{\lambda}\to
D_{-\lambda/2}
{\raise 0.5pt \hbox{$\scriptstyle\circ$}} d_{\lambda/2}.
\end{equation}
The $s\leftrightarrow t$ counterpart of non--relativistic
expansions (\ref{nrctransf})--(\ref{nrctrimp}) with parameter
$\kappa=-\epsilon_{1}$ is in turn a new transformation
we denote by ${\rm C}_{1}$,
\begin{equation}
\begin{array}{ccc}
\hbox{expansions}\qquad\hfill
&
&{\rm C}_{1}
\\
\cr
\widetilde{x}=
\displaystyle\frac{x}{1-\kappa t}\hfill
&
&\widetilde{x}=
\displaystyle\frac{x}{1+\epsilon_{1}s}\hfill
\\
\widetilde{t}=
\displaystyle\frac{t}{1-\kappa t}\hfill
&\Longrightarrow\quad\hfill
&\widetilde{t}=t+
\displaystyle\frac{\epsilon_{1}x^2}{2(1+\epsilon_{1}s)}
\hfill
\\
\widetilde{s}=s-
\displaystyle\frac{\kappa x^2}{2(1-\kappa t)}
\qquad\hfill
&
&\widetilde{s}=
\displaystyle\frac{s}{1+\epsilon_{1} s}\hfill
\\
\end{array}.
\label{expC1}
\end{equation}
The infinitesimal version of the new transformation is
\begin{equation}
X_{8}=\frac{x^2}{2}{\partial}_{t}-xs{\partial}_{x}-s^2{\partial}_{s}.
\label{X8}
\end{equation}
Calculating the Lie brackets of (\ref{X8}) with the other
infinitesimal transformations, we get one more
vectorfield. In fact, the bracket of (\ref{X8}) with
the generator of
infinitesimal boosts, $t{\partial}_{x}-x {\partial}_{s}$, yields
\begin{equation}
X_{9}
=
xt{\partial}_{t}+\big(\frac{x^2}{2}-ts\big){\partial}_{x}+xs{\partial}_{s}.
\label{X9}
\end{equation}
Collecting our results, our symmetry generators read
\begin{equation}
\begin{array}{cccc}
X_{0}\hfill
&=
&{\partial}_{t}\hfill
&\hbox{time translation}\hfill
\\
X_{1}\hfill
&=
&-{\partial}_{x}\hfill
&\hbox{space translation}\hfill
\\
X_{2}\hfill
&=
&-{\partial}_{s}\hfill
&\hbox{vertical translation}\hfill
\\
X_{3}\hfill
&=
&t{\partial}_{x}-x{\partial}_{s}\hfill
&\hbox{galilean boost}\hfill
\\
X_{4}\hfill
&=
&t{\partial}_{t}+\frac{x}{2}{\partial}_{x}\hfill
&\hbox{non-relat. dilatation}\hfill
\\
X_{5}\hfill
&=
&t^2{\partial}_{t}+xt{\partial}_{x}-\frac{x^2}{2}{\partial}_{s}\hfill
&\hbox{expansion}\hfill
\\
X_{6}\hfill
&=
&t{\partial}_{t}-s{\partial}_{s}\hfill
&\hbox{time dilation}\hfill
\\
X_{7}\hfill
&=
&x{\partial}_{t}-s{\partial}_{x}\hfill
&\hbox{``antiboost''}\hfill
\\
X_{8}\hfill
&=
&\frac{x^2}{2}{\partial}_{t}-xs{\partial}_{x}-s^2{\partial}_{s}\hfill
&{{\rm C}_{1}}\hfill
\\
X_{9}\hfill
&=
&xt{\partial}_{t}+\big(\frac{x^2}{2}-ts\big){\partial}_{x}+xs{\partial}_{s}\qquad
\hfill
&{\rm C}_{2}\hfill
\\
\end{array}
\label{Bgenerators}
\end{equation}
The Lie brackets of these vector fields are seen to satisfy the
same algebra as the conserved quantities in (\ref{Poissonbrackets}).
The vectorfields $X_8$ and
$X_{9}$ will be shown below in particular to generate
the two additional conserved quantites ${\rm C}_{1}$ and
${\rm C}_2$ in Eq.
(\ref{newrelconst}).
Note that the algebra (\ref{Bgenerators}) is manifestly invariant
w. r. t. the interchange $t\longleftrightarrow s$.
The vector field $X_{9}$ is itself invariant;
this is the reason why we could not find it by the
``interchange--trick''.
The extended manifold $M$ above has
already been met before.
In their ``Kaluza-Klein-type'' framework
for non-relativistic physics in $d+1$ dimension,
Duval et al. \cite{barg} indeed consider
a $(d+1,1)$--dimensional Lorentz manifold
$\big(M, g_{\mu\nu}\big)$, endowed with a
covariantly constant lightlike ``vertical'' vector $\xi=(\xi^\mu)$
they call ``Bargmann space''.
The quotient of $M$ by the flow of $\xi$ is a non-relativistic
space-time denoted by $Q$.
In the application we have in mind, $M$ is simply
$3$-dimensional Minkowski space, with
the usual coordinates $x_0, x, y$ and metric
$-(dx^{0})^2+dx^2+dy^2$.
Introducing the light-cone coordinates
\begin{equation}
t=\frac{1}{\sqrt{2}}\big(y-x^{0}\big),
\qquad
s=\frac{1}{\sqrt{2}}\big(y+x^{0}\big),
\end{equation}
the Minkowskian metric reads
$dx^2+2dtds$. Then $\xi=\partial_s$
is indeed lightlike and covariantly constant.
All [infinitesimal] conformal transformations of Minkowski space
form the conformal algebra ${\rm o}(3,2)$.
Now, as shown in Appendix A, the $X_{i}$
found above provide just another basis
of this same algebra.
\goodbreak
\section{Conformal geometry}
The action of the orthogonal group ${\rm O}(3,2)$ on
$3$-dimensional Minkowski space is the best described as follows.
Consider the natural action of ${\rm O}(3,2)$ on ${\rm R\hspace*{-2ex} I\hspace*{1.2ex}}^{3,2}$
by matrix multiplication. A vector in
${\rm R\hspace*{-2ex} I\hspace*{1.2ex}}^{3,2}$ can be written as
\begin{equation}
Y=\pmatrix{y\cr a\cr b\cr},
{\qquad\hbox{where}\qquad}
y=\pmatrix{x\cr t\cr s}\in{\rm R\hspace*{-2ex} I\hspace*{1.2ex}}^{2,1},
\;
a,\, b\in{\rm R\hspace*{-2ex} I\hspace*{1.2ex}}.
\end{equation}
The vector space ${\rm R\hspace*{-2ex} I\hspace*{1.2ex}}^{3,2}$ carries the quadratic form
$
{\bar Y}Y={\bar y}y+2ab,
$
where ${\bar y}y$ means
$
{\bar y}y=x^2+2ts,
$
so that $\bar{Y}$ is represented by the row-vector
$(\bar{y}, b, a)$ where
$\bar{y}=(x, s, t)$.
$(2+1)$-dimensional Minkowski space,
$M={\rm R\hspace*{-2ex} I\hspace*{1.2ex}}^{2,1}$, can be mapped into the isotropic
cone (quadric) ${\cal Q}$ in ${\rm R\hspace*{-2ex} I\hspace*{1.2ex}}^{3,2}$, as
\begin{equation}
y\mapsto
\pmatrix{y\cr1\cr-{\smallover 1/2}{\bar y}y\cr}.
\end{equation}
Projecting onto the real projective space $P{\cal Q}$, we
identify $M$ with
those generators in the null-cone in ${\rm R\hspace*{-2ex} I\hspace*{1.2ex}}^{3,2}$.
The manifold $P{\cal Q}$ is invariant with respect to the action
of ${\rm O}(3,2)$.
Let us first consider infinitesimal actions.
An ${\rm o}(3,2)$ matrix can be
written as
\begin{equation}
\pmatrix{
\Lambda &V&W\cr
-{\bar W}&-\lambda &0\cr
-{\bar V}&0 &\lambda\cr
},
\qquad
\Lambda\in\, {\rm o}(2,1),\,V,\,W\in{\rm R\hspace*{-2ex} I\hspace*{1.2ex}}^{2,1},
\;\lambda\in{\rm R\hspace*{-2ex} I\hspace*{1.2ex}}.
\label{o32matrix}
\end{equation}
The matrix action of ${\rm o}(3,2)$ on ${\rm R\hspace*{-2ex} I\hspace*{1.2ex}}^{3,2}$ yields
the action on Bargmann space
\begin{equation}
\Lambda y+V-\2W{\bar y}y+({\bar W}y+\lambda)y.
\label{confalgebraaction}
\end{equation}
In particular, $V$ represents infinitesimal translations.
Observe now that the
covariantly constant null vector $\xi$ is also the
generator of vertical translations,
\begin{equation}
{\hat\xi}=\pmatrix{
0&\xi&0\cr
0&0&0\cr
-\bar{\xi}&0&0\cr
}.
\label{ximatrix}
\end{equation}
The Schr\"odinger algebra is identified as
those vectorfields which commute with the ``vertical vector'',
\begin{equation}
[Z,{\hat\xi}]=0.
\end{equation}
This yields the constraints
$\Lambda\xi=-\lambda\xi$
and
$W=\kappa\xi$,
$\kappa\in{\rm R\hspace*{-2ex} I\hspace*{1.2ex}}$. It follows that
\begin{equation}
Z=\pmatrix{
0
&\beta &0&\gamma&0
\cr
0&\lambda &0&\tau &0
\cr
-{{\beta}}&0&-\lambda&\eta&\kappa\cr
0&-\kappa&0&-\lambda&0\cr
-{\gamma}&-\eta&-\tau&0&\lambda\cr
},
\qquad
\beta,\gamma, \lambda,\kappa,\eta\in{\rm R\hspace*{-2ex} I\hspace*{1.2ex}}.
\end{equation}
This is the {\it extended Schr\"odinger algebra}, with
\vskip1mm
$\bullet$ ${\beta}$ representing Galilei boosts,
$\bullet$ ${\bf \gamma}$
space translations,
$\bullet$ $\tau$ time translations,
$\bullet$ $\lambda$ non-relativistic dilatations,
$\bullet$ $\kappa$ expansions,
$\bullet$ $\eta$ translations in the vertical direction.
\vskip1mm
Using (\ref{confalgebraaction}),
we recover the infinitesimal action of the (extended)
Schr\"o\-din\-ger
algebra on $M$ \cite{barg}.
Note that the {\it relativistic}
dilation invariance [with all directions dilated by the same factor], is
{\it broken} by the reduction:
only doubly time-dilated combinations project to Bargmann space.
Those parameterized by
$\beta, \gamma, \tau,\eta\in{\rm R\hspace*{-2ex} I\hspace*{1.2ex}}$ are isometries
and are recognized as the generators of the {\it extended Galilei}
(or Bargmann) group.
\goodbreak
Let us now identify the unusual generators.
``Antiboosts'' and dilatations of time alone belong
to the upper--left ${\rm o}(2,1)$ corner $\Lambda$ of
the ${\rm o}(3,2)$ matrix,
(\ref{o32matrix})
\begin{equation}
\begin{array}{cc}
\Lambda=\qquad\hfill
&\left\{\begin{array}{cc}
\left(\begin{array}{ccc}
0&0&-\alpha
\\
\alpha&0&0
\\
0&0&0
\end{array}\right)
\qquad\hfill
&\hbox{antiboost}\hfill
\\\cr
\left(\begin{array}{ccc}
0&0&0
\\
0&d&0
\\
0&0&-d
\end{array}\right)
\qquad\hfill
&\hbox{dilation of time alone}\hfill
\\
\end{array}\right.
\end{array}
\end{equation}
Augmented with the extended Galilei
algebra, the Poincar\'e algebra is obtained.
In the same spirit, the two remaining (relativistic)
conformal transformations ${\rm C}_{1}$ and ${\rm C}_{2}$
correspond to chosing
$
W_{1}=\left(\begin{array}{c}
0\\
1\\
0\\
\end{array}\right)
$
and
$
W_{2}=\left(\begin{array}{c}
1\\
0\\
0\\
\end{array}\right),
$
respectively.
The generated group, found by exponentiating, is the conformal
group
${\rm SO}(3,2)$.
The Schr\"odinger group is recovered as those
transformations which commute with the
1-parameter subgroup generated by ${\hat\xi}$.
It acts on $M$ according to
in the standard way which plainly
project to ``ordinary'' space-time and span there
the (non--extended) Schr\"o\-din\-ger group consistently with
(\ref{Galtr}) and (\ref{nrctransf}).
The transformations which do {\it not} preserve $\xi$ are
\begin{equation}
\begin{array}{cc}
\begin{array}{ccc}
\widetilde{x}&=\hfill &x\\
\widetilde{t}&=\hfill &e^{\delta}t
\\
\widetilde{s}&=\hfill &e^{-\delta}s
\end{array}\hfill
&\hbox{time dilation}\hfill
\\
\cr
\begin{array}{ccc}
\widetilde{x}&=\hfill &x-\alpha s
\\
\widetilde{t}&=\hfill &t+\alpha x-\frac{1}{2}\alpha^2s
\\
\widetilde{s}&=\hfill &s
\\
\end{array}\hfill
&\hbox{``antiboost''}\hfill
\\
\cr
\begin{array}{ccc}
\widetilde{x}&=\hfill &
\displaystyle\frac{x}{1+\epsilon_{1}s}
\\
\widetilde{t}&=\hfill &t+
{\smallover 1/2}\displaystyle\frac{\epsilon_{1}x^2}{1+\epsilon_{1}s}
\\
\widetilde{s}&=\hfill &
\displaystyle\frac{s}{1+\epsilon_{1}s}
\\
\end{array}
\hfill
&{\rm C}_{1}
\\
\cr
\begin{array}{ccc}
\widetilde{x}&=\hfill &
\displaystyle{
\frac{x-\epsilon_{2}\big(\2x^2+ts\big)}
{(1-{\smallover 1/2}\epsilon_{2}x)^2+{\smallover 1/2}\epsilon_{2}^2ts}}
\\
\widetilde{t}&=\hfill &
\displaystyle{
\frac{t}{(1-{\smallover 1/2}\epsilon_{2}x)^2+{\smallover 1/2}\epsilon_{2}^2ts}}
\\
\widetilde{s}&=\hfill &
\displaystyle{
\frac{s}{(1-{\smallover 1/2}\epsilon_{2}x)^2+{\smallover 1/2}\epsilon_{2}^2ts}}
\\
\end{array}
\qquad\hfill
&{\rm C}_{2}
\\
\end{array}
\label{Cgroupaction}
\end{equation}
Our transformations are indeed conformal since they satisfy
$
f^*g_{\mu\nu}=\Omega^2g_{\mu\nu}
$
see Appendix A.
Note that the interchange
$s\leftrightarrow t$ is also an isometry and
carries the group {\rm SO}$(3,2)$ into another component of the
conformal group ${\rm O}(3,2)$.
\goodbreak
\section{Projecting to ordinary space--time}
As we said already, the quotient $Q$ is $1+1$-dimensional
``ordinary'' spacetime,
labeled by $x$ (position) and $t$ (time).
The projection
$M\to Q$ means simply ``forgetting'' the vertical coordinate
$s$~:
$\left(\begin{array}{c}
x\\ t\\ s
\end{array}\right)
\to
\left(\begin{array}{c}
x\\ t
\end{array}\right).
$
Next, we wish to relate the fields on extended and on
ordinary space, respectively.
Let us recall how this is done usually \cite{barg}.
Let $\psi$ denote a complex field
on $M$. Then, requiring the field to be equivariant,
\begin{equation}
\xi^\mu{\partial}_{\mu}\psi=i\psi,
\label{oldequiv}
\end{equation}
allows us to reduce $\psi$
from Bargmann space to one on ordinary space-time as
$\Psi(x,t)=e^{-is}\psi(x,t,s)
$
\cite{barg}. Writing $\psi=\rho^{1/2}e^{i\theta}$,
(\ref{oldequiv}) reads
$
\xi^\mu\partial_\mu\rho=0
$
and
$
\xi^\mu\partial_\mu\theta=1.
$
In light--cone coordinates of Minkowski case in particular,
these conditions imply that
\begin{equation}
R(x,t)=\rho(x,t,s),
\qquad
\Theta(x,t)=\theta(x,t,s)-s.
\label{modequiv}
\end{equation}
are well--defined fields on $Q$. These formul{\ae}
(also referred to as equivariance) allows us to relate
equivariant
fields on extended space to fields on ordinary space.
Let us now consider a diffeomorphism
\begin{eqnarray}
f(x,t,s)\equiv
\left(\begin{array}{c}
\widetilde{x}
\\
\widetilde{t}
\\
\widetilde{s}
\end{array}\right)
=
\left(\begin{array}{c}
g(x , t, s)
\\
h(x , t, s)
\\
k(x, t, s)
\end{array}\right)
\label{difofM}
\end{eqnarray}
of $M$. How can we
project this to ordinary space--time~?
In the particular case when the mapping preserves $\xi$,
the entire fibre goes into the same fibre and the result
projects to a well--defined diffeomorphism of ordinary
space-time.
In fact,
$
\widetilde{x}=g(x,t)
$,
$
\widetilde{t}=h(x,t)
$,
$\widetilde{s}\equiv k(x,t,s)=s+K(x,t)
$,
so that we can define the projected map
$F(x,t)=\left(\begin{array}{c}
x^\star
\\
t^\star\end{array}\right)
$
by setting
$
\begin{array}{c}
x^\star=\widetilde{x}=g(x,t)
$
and
$
t^\star=\widetilde{t}=h(x,t).
\end{array}
$
As a bonus, we also get
the usual transformation
rule of the phase (consistent with the equivariance),
$
\begin{array}{c}
\Theta^\star(x,t)=\Theta(x^\star,t^\star)+K(x,t).
\end{array}
$
If, however, $f$ does not preserve the fibres, this construction
does
not work since the coordinates $\widetilde{x}$ and
$\widetilde{t}$
now depend on $s$. Hence the need of generalizing the
construction
based on equivariance. Forgetting momentarily about
$\rho$, we only
consider the phase, $\theta$.
Our clue is to observe that if $\theta$ is equivariant,
(\ref{modequiv}),
then $s=-\Theta(x,t)$ is solution of the equation
$\theta(x,t, s)=0$, i. e.,
\begin{equation}
\theta\big(x,t,-\Theta(x,t)\big)=0.
\label{newequiv}
\end{equation}
This condition is, however, meaningful without any assumption of
equivariance and associates implicitly a function
$\Theta(x,t)$ to each
$x$, $t$ {\it and} field $\theta$. Conversely, to any
$x$, $t$ and $\Theta(x,t)$ Eq. (\ref{modequiv}) associates an
(equivariant) field $\theta(x,t,s)$ on $M$.
Let us recall that
our extended ``Bargmann'' space $M$ is a fibre bundle over
ordinary space-time $Q$, with fibre ${\rm R\hspace*{-2ex} I\hspace*{1.2ex}}$. Then
$\Theta$ corresponds to a section $Q\to M$ of this bundle.
Condition (\ref{newequiv}) requires the existence of a section
$\Theta(x,t)$ along which the phase field $\theta$ vanishes.
A diffeomorphism $f$
of $M$ acts naturally on $\theta$, namely as
\begin{equation}
\widetilde{\theta}=f^\star\theta.
\label{actionontheta}
\end{equation}
We can define therefore $\Theta^\star$ as the solution
of the equation
\begin{equation}
\widetilde\theta(x,t,-\Theta^\star(x,t))=0.
\label{projcond}
\end{equation}
This implicit equation (assumed to admit a unique
solution) associates a $\Theta^\star$ to $x$, $t$ and
$\theta$.
Let us stress that
$\widetilde{\theta}$ is not in general equivariant even
if $\theta$ is equivariant, unless $f$ preserves the fibres.
Thus,
starting with $\Theta(x,t)$ we lift it first to $M$ as an
equivariant field $\theta(x,t,s)=\Theta(x,t)+s$ on $M$; to
which a
well--defined $\Theta^\star$ (function of
$x$, $t$ and $\Theta$) is associated by (\ref{projcond}).
Having defined
$\Theta^\star$,
the diffeomorphism $f$ of $M$
can be projected to $Q$ in a ``field--dependent way'' by
restricting
$\widetilde{f}$ to the section $s=-\Theta^*$. In
coordinates,
$F(x,t)=
\left(\begin{array}{c}
x^\star
\\
t^\star
\\
\end{array}\right)
$, where
\begin{equation}
\begin{array}{c}
x^\star=g\big(x,t,-\Theta^\star(x,t)\big),
\\
t^\star=h\big(x,t,-\Theta^\star(x,t)\big),
\\
\Theta(x^\star,t^\star)=
-k\big(x,t,-\Theta^\star(x,t)\big).\hfill
\end{array}
\label{actiononQ}
\end{equation}
The last line here requires to express
$\Theta^\star$ by inverting the function $k$ and
reinserting the result
into the two first lines.
It also implements the transformation
on the ``phase'', $\Theta$.
Let us stress that these formul{\ae} are implicit~:
$x^\star$ and $t^\star$ can not be defined without defining
${\Theta}^\star $, which itself involves $x^\star$ and $t^\star$.
In the equivariant case, the procedure is
plainly consistent with the previous formulae.
In the non-fiber-preserving case
it yields the ``field--dependent
diffeomorphisms'' considered by Bazeia and Jackiw \cite{BJ}.
For ``antiboosts'', for example, we get from
(\ref{BJBtransf})
\begin{equation}
\begin{array}{cc}
&{x}^\star=
x+\alpha{\Theta}^\star (x, t),
\hfill
\\
&{t}^\star=t+\alpha x+\frac{1}{2}\alpha^2
{\Theta}^\star (x, t)
\\
&{\Theta}^\star(x, t)=
\Theta(x^\star, t^\star),
\hfill
\end{array}
\label{ABdef}
\end{equation}
which is equivalent to the definition (\ref{BJtransf}).
Time dilations work similarly.
The formulae valid for the
two relativistic conformal transformations,
${\rm C}_{1}$ and ${\rm C}_{2}$ above, is presented in
Appendix B (\ref{rctransf}).
The formula for ${\rm C}_{1}$ is consistent with (\ref{expC1});
that for ${\rm C}_{2}$ is a new result.
So far, we only studied how to act on $\Theta$~:
(\ref{actiononQ}) only involves the phase but not
the density.
Turning to this problem, let us {\it posit}
\begin{equation}
R(x,t)=
\rho\big(x,t,-\Theta(x,t)\big)
{\partial}_{s}\theta\big(x,t,-\Theta(x,t)\big),
\label{Rdef}
\end{equation}
where $\Theta$ is defined by (\ref{newequiv}).
$R(x,t)$ is a well--defined function of $x$ and $t$.
Let us insist that (\ref {Rdef}) is again ``field--dependent'' in that it
also depends on $\theta$, except when
$\theta$
is equivariant, when it reduces to (\ref{modequiv}).
Conversely, if $R(x,t)$ is any field on $Q$,
$\rho(x,t,s)=R(x,t)$ can
obviously be viewed as
(an equivariant) function on extended space.
Let us henceforth consider a conformal transformation $f$
of $M$ $f^\star g_{\mu\nu}=\Omega^2g_{\mu\nu}$ and let $\rho$
be a (possibly not equivariant) field on $M$.
$f$ acts naturally on $\rho$ as
\begin{equation}
\rho\to\widetilde{\rho}=\Omega f^*\rho.\hfill
\label{Brhoimpl}
\end{equation}
Hence
\begin{equation}
R^\star(x,t)=
\widetilde{\rho}\big(x,t,-\Theta^\star(x,t)\big)
{\partial}_{s}\widetilde{\theta}\big(x,t,-\Theta^\star(x,t)\big).
\label{Rstardef}
\end{equation}
Using the definition (\ref{Rdef}) of $R$, this is also
written as
\begin{equation}
R^\star(x,t)=
\Omega(x,t,-\Theta^\star)\,
\frac{{\partial}_{s}\widetilde{\theta}
\big(x,t,-{\Theta}^\star(x,t)\big)}
{{\partial}_{\widetilde{s}}\theta\big(x^\star,t^\star,
-{\Theta}(x^\star,t^\star)\big)}
R({x}^\star,{t}^\star).
\label{Rimp}
\end{equation}
(If the field $\theta$ is equivariant,
the denominator is equal to $1$).
On the other hand, one can show in general that
\begin{equation}
\frac{{\partial}_{s}\widetilde{\theta}
\big(x,t,-{\Theta}^\star(x,t)\big)}
{{\partial}_{\widetilde{s}}\theta\big(x^\star,t^\star,
-{\Theta}(x^\star,t^\star)\big)}
=
\frac{\widetilde{J}\big(x,t,-{\Theta}^\star(x,t)\big)}
{J^\star(x,t)},
\end{equation}
where ${J}^*$ and
$\widetilde{J}$
are the Jacobians on ordinary and on the extended space
respectively,
\begin{equation}
\begin{array}{cc}
J^\star
={\rm det}\left(
\displaystyle{\frac{{\partial}\big(x^\star\big)^\alpha}
{{\partial} x^\beta}}
\right)\qquad
\hfill
&\widetilde{J}
={\rm det}\left(
\displaystyle{\frac{{\partial}\widetilde{x}^\mu}
{{\partial} x^\nu}}\right),
\end{array}
\label{jaconB}
\end{equation}
($\alpha, \beta=x, t$ and $\mu,\nu=x, t, s$).
Eq. (\ref{Rimp}) can therefore be rewritten as
\begin{equation}
R^\star(x,t)
=
\Omega\big(x,t,-{\Theta}^\star(x,t)\big)\times
\frac{\widetilde{J}\big(x,t,-{\Theta}^\star(x,t)\big)}
{J^\star(x,t)}\,R(x^\star,t^\star).
\label{imponR2}
\end{equation}
For a conformal transformation $\widetilde{J}=\pm \Omega^3$,
the
sign depending on the mapping being orientation--preserving
or not.
If the transformation $f$ preserves $\xi$, $\Omega$ is a
function
of $t$--alone cf. (\ref{conffactor}) in Appendix A. Then
$\widetilde{\theta}$ is
again equivariant and our formula reduces to the standard
expression
\begin{equation}
R^\star(x,t)
=
\Omega(t)\,
R(x^\star,t^\star),
\label{imponR3}
\end{equation}
cf. (\ref{nrctrimp}).
For an isometry, $\Omega=1$, so that
(\ref{imponR2}) reduces to
\begin{equation}
R^\star(x,t)
=
\frac{R(x^\star,t^\star)}
{J^\star(x,t)}.
\end{equation}
For time dilations
and ``antiboosts'',
the formul{\ae} of Bazeia and Jackiw in \cite{BJ},
(our (\ref{BJconst})), are recovered.
For the relativistic conformal transformations
${\rm C}_{1}$ and ${\rm C}_{2}$,
we find some complicated expressions
(\ref{imppr}), presented in Appendix B.
Our formulae allow to implement any
isometry of $M$, not only those in
the connected component of the Poincar\'e
group. Let us consider, for example, the interchange
\begin{equation}
t\longleftrightarrow s,
\end{equation}
which is a non--fiber-preserving isometry. It acts on the
fields defined on $M$ in the natural way.
For fields on $Q$, we get the ``field-dependent action''
\begin{equation}
\begin{array}{c}
x^\star=x,
\ccr
t^\star=-\Theta^\star(x,t),
\ccr
\Theta\big(x,-\Theta^\star\big)+t=0,
\ccr
R^\star(x,t)=R\big(x,-\Theta^\star\big){\partial}_{t}
\Theta(x,-\Theta^\star)
=\displaystyle{
\frac{R\big(x,-\Theta^\star\big)}{{\partial}_{t}\Theta^\star(x,t)}}.
\end{array}
\label{stimpl}
\end{equation}
This formula is so much implicit that we can not go farther unless
$\Theta$ is given explicitly. It is nevertheless a ``field--dependent
symmetry''.
The weak condition (\ref{projcond}) can hence accomodate the
$t\leftrightarrow s$ symmetry.
This is in sharp contrast with the equivariance condition
(\ref{oldequiv}) which manifestly breaks it.
Note also that our formul{\ae} for implementing
the conformal transformations on the fields are consistent
with the
interchange symmetry $t\leftrightarrow s$, followed by the rule of replacing
$s$ with $-\Theta^\star$.
When applied to a Schr\"odinger transformation, it yields
its non--$\xi$--preserving counterpart.
\section{Physics on extended space}
So far the ``Bargmann space'' $M$ was only used as a geometric
arena for linearizing the action of the conformal group.
Now we show how to lift the {\it physics} to $M$.
Generalising our
previous theory, let $M$ be
$(d+1,1)$--dimensional Lorentz manifold
$\big(M, g_{\mu\nu}\big)$ endowed with a
covariantly constant lightlike vector $\xi=(\xi^\mu)$.
Such a manifold admits a preferred coordinates
$\vec{x}, t, s$ in which the metric is
\begin{equation}
g_{ij}(\vec{x},t)dx^idx^j+2dt\big[ds+\vec{A}\cdot d\vec{x}\big]
-2U(\vec{x},t)dt^2,
\label{BRINK}
\end{equation}
where $g_{ij}$ is a metric on $d$--dimensional
``transverse space''
and $\vec{A}$ and $U$ are vector and scalar potentials,
respectively \cite{BRINK}, \cite{barg}.
\subsection{Field theory on extended space}
Let $\rho$ and $\theta$ be two real fields on $M$, and
let us consider the field theory described by the
action
\begin{equation}
\begin{array}{cc}
S=&S_{0}+S_{p}=\hfill
\ccr
&\displaystyle{
\int-\frac{1}{2}\left(\rho\nabla_\mu\theta\nabla^\mu\theta
\right)
\sqrt{-g}\,d^3x
-\int
V(\rho)\sqrt{-g}\,d^3x},
\hfill
\end{array}
\label{barlag}
\end{equation}
where $\nabla_{\mu}$ is the covariant derivative associated
with the
metric of $M$.
The Euler-Lagrange equations read
\begin{equation}
\nabla_\mu(\rho\nabla^\mu\theta)=0,
\qquad
{\smallover 1/2}\nabla_\mu\theta\nabla^\mu\theta=-\frac{dV}{d\rho}.
\label{bareq}
\end{equation}
When the fields are required to be also equivariant,
(\ref{modequiv}), then, for the projected variables
$\Theta$ and $\rho$, the equations of motion (\ref{bareq})
reduce to
those of Bazeia and Jackiw in Ref. \cite{BJ},
Eqn. (\ref{eqmotion}) above. (Working with a general
Bargmann space \cite{barg} would allow us to describe our
fluid system in
an external electromagnetic field).
Equivariance is a too strong condition, though.
For specific potentials, the weaker conditions (\ref{newequiv})
and
(\ref{Rimp}), i.e.
\begin{equation}
\begin{array}{c}
\theta\big(x,t,-\Theta(x,t)\big)=0,\hfill
\\
R(x,t)=
\rho\big(x,t,-\Theta(x,t)\big)
{\partial}_{s}\theta\big(x,t,-\Theta(x,t)\big)\hfill
\end{array}
\label{weakc}
\end{equation}
may still work. Expressing ${\partial}_{\alpha}\Theta$ by deriving
the defining relation (\ref{newequiv}) one finds using the
Euler--Lagrange equations (\ref{bareq}) that $R$ and $\Theta$
satisfy the Bazeia--Jackiw (\ref{eqmotion}) {\it provided}
$V(\rho)$ is the membrane potential $V(\rho)=c/\rho$. This
is hence
the only potential consistent with (\ref{newequiv}).
In sharp contrast with equivariance,
our new condition does not impose any restriction to the fields.
Let us consider, for example, the action
(\ref{barlag}) on $(2+1)$ dimensional Minkowski space with
$V=0$ and chose
\begin{equation}
\rho=\sqrt{R}
\qquad\hbox{and}\qquad
\theta=\sqrt{R}\,\sin(\Theta+s).
\end{equation}
The corresponding field on Bargmann space,
$\psi={R}^{1/4}e^{i\sqrt{R}\,\sin(\Theta+s)}$, is not equivariant.
This Ansatz satisfies however our conditions (\ref{weakc}),
as anticipated by
the notations. It projects to (\ref{BJ})
with its large symmetry.
\goodbreak
\subsection{Symmetries}
The Kaluza-Klein type framework is particularly
convenient for studying the symmetries.
Let us indeed consider a conformal diffeomorphism
$f(x,t,s)$ of the Bargmann metric.
It is easy to see, along the lines indicated in
Refs. \cite{barg}, that implementing $f$ on the fields as
\begin{equation}
\begin{array}{c}
\theta\to\widetilde{\theta}=f^*\theta,\hfill
\\
\rho\to\widetilde{\rho}=\Omega f^*\rho,\hfill
\end{array}
\label{impl}
\end{equation}
the ``free'' action (\ref{barlag}) is left
invariant by all conformal transformations of $M$.
This is explained by the absence of any mass term in
(\ref{barlag}).
Equivalently, the transformed fields are seen to satisfy
the equations of motion
\begin{equation}
\nabla_\mu(\widetilde\rho\,\nabla^\mu\widetilde\theta)=0,
\qquad
\nabla_\mu\widetilde\theta\,\nabla^\mu\widetilde\theta=0.
\label{eqonB}
\end{equation}
It is worth noting that
unfolding to extended space converted the
up--to-surface--term invariant system (\ref{BJ}) into a
strictly invariant one.
\goodbreak
We can now derive once again the symmetries
starting from the extended space. Let us first consider the
free case.
Differentiating the defining relations
(\ref{projcond}) and (\ref{Rstardef})
we find, using the equations of motion (\ref {eqonB}) on $M$,
that $R^\star$ and $\Theta^\star$
satisfy the free equations of motion in ordinary space,
\begin{eqnarray*}
\partial_{t}R^\star+\partial_{x}\big(
R^\star\partial_{x}\Theta^\star\big)=0,
\qquad
\partial_t\Theta^\star+\frac{1}{2}
\big(\partial_x\Theta^\star\big)^2=0.
\end{eqnarray*}
Alternatively, we check readily that
\begin{eqnarray*}
dxdtR^\star(x,t)\Big[{\partial}_{t}\Theta^\star(x,t)
+\frac{1}{2}\big({\partial}_{x}\Theta^\star(x,t)\big)^2\Big]
=\hfill
\\
dx^\star dt^\star R(x^\star, t^\star)
\Big[{\partial}_{t^\star}\Theta(x^\star, t^\star)
+\frac{1}{2}\big({\partial}_{x^\star}\Theta(x^\star,t^\star)\big)^2
\Big].
\hfill
\end{eqnarray*}
The free action (\ref{BJ}) is hence invariant~:
each conformal transformation of extended space
projects to a symmetry of the free system.
Restoring the potential term,
the scaling properties imply again that
conformal symmetry on $M$ only allows
$
V=c\rho^3,
$
cf (\ref{confinvpot}).
This potential is, however, inconsistent with the generalized
condition (\ref{newequiv}) unless $c=0$. Then we have the choice~:
if we keep $V=c\rho^3$ and use the
usual equivariance (\ref{oldequiv}), then the non--fiber preserving
part is broken and we are left with a Schr\"odinger symmetry.
If we choose $V=c/\rho$ the conformal symmetry is broken to its
Poincar\'e subgroup from the outset; this survives, however, the
reduction based on the generalised condition (\ref{newequiv}).
In particular, the interchange
$t\leftrightarrow s$,
implemented as in (\ref{impl}) on $\theta$ and $\rho$
(or on $\Theta$ and $R$ as in (\ref{stimpl})) is a symmetry.
\goodbreak
\subsection{Conserved quantities}
On Bargmann space, we have a relativistic theory.
Defining the energy--momentum tensor
as the variational derivative of
the action w. r. t. the metric,
$
{\cal T}_{\mu\nu}=2{\delta S}/{\delta g^{\mu\nu}},
$
we find
\begin{eqnarray}
{\cal T}_{\mu\nu}=-\rho\,\nabla_\mu\theta\nabla_\nu\theta
+\frac{\rho}{2}\,g_{\mu\nu}\nabla_\sigma\theta\nabla^\sigma\theta
+g_{\mu\nu}V(\rho).
\label{Bemom}
\end{eqnarray}
This energy--momentum tensor is symmetric,
${\cal T}_{\mu\nu}={\cal T}_{\nu\mu}$, by construction and also manifestly.
Using the equation of motion (\ref{bareq}),
we see at once that ${\cal T}_{\mu\nu}$ is traceless,
${\cal T}^\mu_{\ \mu}=0$,
precisely when $V=c\rho^3$ i.e., when our theory has the
conformal symmetry.
Finally, ${\cal T}_{\mu\nu}$ is conserved,
\begin{eqnarray}
\nabla_\mu{\cal T}^{\mu\nu}=0,
\end{eqnarray}
as it follows from general covariance
(i. e. from covariance w. r. t. diffeomorphisms \cite{SOUR}),
and also from the eqns. of motion.
Let us assume that the potential is $V(\rho)=c\rho^3$ so that the
system has conformal symmetry.
To any conformal vector field
${X}=\big({X}^\mu\big)$ on $M$,
$L_{{X}}g_{\mu\nu}=\lambda g_{\mu\nu}$,
we can now associate a conserved current \cite{DHP} on
$M$ by contracting the
energy-momentum tensor
\begin{eqnarray}
k^\mu={\cal T}^{\mu}_{\ \nu}X^\nu.
\label{conscur}
\end{eqnarray}
In fact,
$\nabla_{\mu}k^\mu
=
(\nabla_\mu{\cal T}^{\mu}_{\ \nu}){X}^\nu
+
\frac{1}{2}L_{{X}}g_{\mu\nu}{\cal T}^{\mu\nu}=0.
$
The first term here vanishes
beacause ${\cal T}_{\mu\nu}$ is conserved, and the second term
vanishes because ${\cal T}_{\mu\nu}$ is traceless.
Let us assume henceforth that the fields $\rho$ and $\theta$
are also equivariant.
Then the Bargmann--space energy--momentum tensor ${\cal T}_{\mu\nu}$
becomes $s$-independent.
If $X^\mu$ commutes with the vertical vector $\xi^\mu$,
one can construct a conserved current on ordinary space out of
$k^\mu$ as follows \cite{DHP}. $k^\mu$ does not depend on $s$ and
projects therefore into a well-defined current $J^\alpha$ on $Q$,
$(k^\mu)=(J^\alpha, k^s)$.
The projected current is furthermore conserved,
$\nabla_{\alpha}J^\alpha=0$,
because $\xi=\nabla_{s}$ is covariantly constant so that
$\nabla_{s}k^s=0$.
In the general case, however,
the current $k^\mu$ can not be projected in
ordinary space, because it may depend on $s$;
$\nabla_{s}k^s$ may also be non-vanishing.
Our idea is to construct a new current out of $k^\mu$ which does
have the required properties.
Let us restrict in fact $k^\mu$ to a ``section''
$s=-\Theta(x,t)$, i.e., define the Bargmann--space vector
\begin{equation}
j^\mu(x,t)=k^\mu\big(x,t,-\Theta(x,t)\big)
=
\Big({\cal T}^\mu_{\ \nu}X^\nu\Big)
\big(x,t,-\Theta(x,t)\big).
\label{projcur}
\end{equation}
Then
\begin{eqnarray}
\nabla_{\mu}j^\mu
=
\nabla_{\alpha}k^\alpha
-\nabla_{\alpha}\Theta\nabla_{s}k^\alpha
=
-\nabla_{s}\Big(\nabla_{x}\Theta\, k^x
+\nabla_{t}\Theta\, k^t+k^s\Big).
\label{kchange}
\end{eqnarray}
Inserting here the explicit form of $k^\mu$ we find that the
bracketed quantity vanishes due to the equations of motion.
The current $j^\mu$ is therefore conserved on $M$,
$
\nabla_{\mu}j^\mu=0.
$
Let us now define the projected current as
\begin{equation}
J^\alpha(x,t)=
\frac{j^\alpha(x,t)}{\nabla_{s}\theta\big(x,t,-\Theta(x,t)\big)}.
\end{equation}
It can shown using the equations of motion that $J^\alpha$ is a
conserved current on $Q$,
$
\nabla_{\alpha}J^\alpha=0.
$
Integrating the time--component of the projected current
on ordinary space,
\begin{equation}
\int dx J^t
\equiv
\int\!dx\,\frac{{\cal T}_{\mu\nu}}{\nabla_{s}\theta}
X^\mu\xi^\nu
\equiv
\int\! dx\,\frac{{\cal T}_{\mu s}}{\nabla_{s}\theta}X^\mu,
\end{equation}
is hence conserved for any conformal vector $X=(X^\mu)$.
This yields the same conserved quantities as found before.
The Barg\-mann-space energy--momentum tensor is in fact
related to that in ordinary space, (\ref{BJem}), according to
\begin{equation}
\begin{array}{ccccc}
(\nabla_{s}\theta)\, T_{tt}\hfill&=&-{\cal T}^t_{t}\hfill&=
&-{\cal T}_{st},\hfill
\\
(\nabla_{s}\theta)\, T_{tx}\hfill&=&{\cal T}^t_{x}\hfill&=
&{\cal T}_{sx},\hfill
\\
(\nabla_{s}\theta)\, T_{xt}\hfill&=&-{\cal T}^x_{t}\hfill&=
&-{\cal T}_{xt},\hfill
\\
(\nabla_{s}\theta)\, T_{xx}\hfill&=&{\cal T}^x_{x}\hfill&=
&{\cal T}_{xx}.\hfill
\label{emtensors}
\end{array}
\end{equation}
Owing to the extra dimension,
the Bargmann--space energy--momentum tensor admits
the new component ${\cal T}_{ss}$ which, when contracted with the
``vertical'' component $X^s$ of the lifted vector field, yields
the $-C^0$ term in Noether's theorem (\ref{consquant}).
The situation is nicely illustrated by formulae like
(\ref{Ttransc}) of Section 2.
When $X$ is fiber--preserving, we recover the generators
$H, P, B, N, \Delta, K$ in
(\ref{Galconst})
and (\ref{newnrconst}) of the Schr\"odinger algebra.
For the non--fiber--preserving vectors,
we get instead the new conserved quantities $G, D,
{\rm C}_1, {\rm C}_{2}$
in (\ref{BJconst}) and (\ref{newrelconst}).
Interchange, $t\leftrightarrow s$, acts on the Lie algebra of
conserved quantities \cite{Duval}. It carries in particular
the energy to particle density, boosts to ``antiboosts'',
etc.,
as already noted in Section 3.
\section{The symmetries of the Schr\"o\-dinger equation}
We discuss now the (non--linear) Schr\"odinger
equation in $d$ spatial dimensions,
\begin{equation}
i{\partial}_{t}\Psi
=
-\frac{1}{2}\bigtriangleup\Psi
-\frac{{\partial}\overline{V}(\vert\Psi\vert^2)}{{\partial}\Psi^{*}}.
\label{NLS}
\end{equation}
where $\bigtriangleup$ is the $d$-dimensional Laplacian.
When the wave function is
decomposed into module, $R$, and
phase, $\Theta$,
$
\Psi= R^{1/2}e^{i\Theta},
$
Eqn. (\ref{NLS}) becomes indeed (\ref{eqmotion}), with
\begin{equation}
V=\overline{V}+\frac{1}{8}
\frac{({\partial}_{i}R)^2}{R}.
\label{effectivepot}
\end{equation}
A non-vanishing effective potential $V$ is obtained therefore
even for the {\it linear} Schr\"odinger equation
$\overline{V}=0$.
The ``free'' theory
described by ${\cal L}_{0}$ in Eqn. (\ref{BJ}) corresponds hence to
a non--linear Schr\"odinger equation (\ref{NLS}) with
effective potential
$\overline{V}=-\frac{1}{8}
\frac{\nabla_{i}R\nabla^{i}R}{R}$, this latter
canceling the term coming from the hydrodynamical
transcription.
As we show below,
canceling this effective term plays a crucial role.
Let us explain everything from the ``Kaluza-Klein type'' viewpoint.
Generalizing to curved space, let us
consider a complex scalar field $\psi$ on a $d+2$ dimensional
``Brinkmann'' space $M$ (\ref{BRINK}).
Generalizing the flat--space results, we posit the action
\begin{equation}
S=
\int\frac{1}{2}\nabla_{\mu}\psi\,\nabla^{\mu}\bar{\psi}
\,\sqrt{-g}\,d^{d+2}x,
\label{BNLS}
\end{equation}
where $g={\rm det}\big(g_{\mu\nu}\big)$. The associated field equation
is the curved--space massless Klein-Gordon
(i.e., the free wave) equation
\begin{equation}
\nabla_{\mu}\nabla^{\mu}\psi=0.
\label{waveq}
\end{equation}
Equation (\ref{waveq}) is {\it not} in general invariant
w. r. t. conformal transformations of $M$,
$f^\star g_{\mu\nu}=\Omega^2g_{\mu\nu}$, implemented
as
\begin{eqnarray}
\psi\to\widetilde{\psi}=\Omega^{d/2}f^*\psi.
\end{eqnarray}
We explain this in the hydrodynamical transcription.
Decomposing
$\psi$ as $\psi=\sqrt{\rho}\,e^{i\theta}$, the action
(\ref{BNLS}) becomes
\begin{equation}
S=
\int\left(\frac{1}{2}\rho\nabla_{\mu}
\theta\nabla^{\mu}\theta
+
\frac{1}{8}\frac{\nabla_{\mu}\rho\nabla^{\mu}\rho}
{\rho}
\right)\sqrt{-g}\,d^{d+2}x.
\label{Baction}
\end{equation}
The action on the fields is now (\ref{impl}) i.e.
$
\theta\to\widetilde{\theta}=f^*\theta
$,
$
\rho\to\widetilde{\rho}=\Omega^d f^*\rho
$.
As we have seen before,
the first (``kinetic'') term in (\ref{Baction}) is invariant.
The second term is {\it not} invariant.
Let us, however, modify the Lagrangian by adding a term
which involves the
scalar curvature ${\cal R}$ of $M$,
\begin{equation}
S_{\cal R}=
\int\underbrace{\Big[
\frac{1}{2}\nabla_{\mu}\psi\nabla^{\mu}\bar{\psi}
+\frac{d}{8(d+1)}{\cal R}\vert\psi\vert^2
\Big]}_{{\cal L}_{R}}
\sqrt{-g}\,d^{d+2}x.
\label{BNLSmod}
\end{equation}
Then the symmetry-breaking terms will be absorbed by those
which come
from transforming ${\cal R}$, leaving a mere surface term
(see Appendix C).
In conclusion, the conformal symmetry {\it on M}
is restored by the
inclusion of the scalar curvature term as in Eq.
(\ref{BNLSmod}),
see \cite{CCJ}.
Let us stress that this curvature term is only necessary due
to the
presence of the non-linear potential
$\frac{1}{8}\frac{\nabla_{\mu}\rho\nabla^{\mu}\rho}{\rho}$.
Restoring the potential, the conformally invariant action is
\begin{equation}
S_{\bar{V}}=
\int\left[\frac{1}{2}\nabla_{\mu}\psi\,\nabla^{\mu}\bar{\psi}
+\frac{d}{8(d+1)}{\cal R}\vert\psi\vert^2
-\overline{V}(\psi^\star\psi)
\right]\sqrt{-g}\,d^{d+2}x.
\label{BVNLS}
\end{equation}
In Minkowski space ${\cal R}\equiv0$. The curvature- term
must nevertheless be added to the Lagrange density, since the
confor\-mal\-ly--transformed metric has already ${\cal R}\neq0$.
The scaling properties of the Lagrangian imply furthermore that
$\overline{V}(\rho)=c\rho^{1+2/d}$ is the only potential
consistent
with the conformal symmetry ${\rm O}(d+2,2)$.
So far, we have only considered what happens on extended space.
When the theory is reduced to ordinary space-time, some of the
symmetries will be lost, however. We explain this when $M$ is
$(2+1)$--dimensional Minkowski space and for the linear
Schr\"odinger
equation $\overline{V}=0$.
Firstly, the full conformal group (or its Poincar\'e subgroup)
can only be projected
to a (field--depen\-dent) action on ordinary space-time
using (\ref{newequiv}) and (\ref{projcond}).
However,
the extended--space model only reduces to
one of the Bazeia-Jackiw form (\ref{BJ}) on $Q$ when the
potential is
$V(\rho)=c/\rho$.
The effective potential in (\ref{Baction})
is manifestly not of this form, though.
The weak condition (\ref{newequiv}) is
hence inconsistent with the Schr\"odinger equation and has
therefore
to be discarded.
Under the assumption of equivariance instead,
Eq. (\ref{modequiv}), the wave equation (\ref{waveq})
on Minkowski space reduces,
for $\Psi(x,t)=e^{-is}\psi(x,t,s)$, to the free
Schr\"o\-din\-ger equation
\begin{equation}
i{\partial}_{t}\Psi+\frac{1}{2}{\partial}_{x}^{2}\Psi=0.
\end{equation}
In terms of $R(x,t)$ and $\Theta$ where
$\Psi=\sqrt{R}e^{i\Theta}$,
this equation becomes
\begin{equation}
\begin{array}{c}
{\partial}_{t}R+{\partial}_{x}\big(R{\partial}_{x}\Theta\big)=0,\hfill
\\
{\partial}_t\Theta+\frac{1}{2}({\partial}_x\Theta)^2=
-\frac{1}{8}\frac{({\partial}_{x}R)^2}{R^2}+\frac{{\partial}_{x}^2R}{4R}.
\hfill
\label{schroddec}
\end{array}
\end{equation}
(Eqn. (\ref{schroddec}) does not contradict
(\ref{eqmotion}), since now $V=V(R,{\partial}_{x}R)$).
As explained in Section 5, usual equivariance only allows
the Schr\"odinger subgroup to project~: the ``truly
relativistic''
generators $G$ and $D$, (i.e., the antiboosts and the
time dilations)
as well as conformal generators ${\rm C}_{1}$ and
${\rm C}_{2}$
are hence broken by the reduction, leaving
us with the mere Schr\"odinger symmetry \cite{JHN},
\cite{barg}.
This latter is furthermore consistent with the potential
$\overline{V}(\rho)=c\rho^{1+2/d}$.
The conserved quantities can be determined as indicated above.
For the linear Schr\"odinger equation in $(1+1)$ dimensions,
for example,
the conserved energy-momentum tensor (\ref{Schemt})
in Appendix C
allows to calculate the conserved quantities.
{\it On extended space} all conformal transformations are
symmetries,
and (\ref{conscur}) associates a conserved current
$k^\mu(,x,t,s)$,
on $M$, $\nabla_{\mu}k^\mu=0$, to each conformal generator.
Its restriction to
the section $s=-\Theta(x,t)$,
$
j^\mu(x,t)
$
in (\ref{projcur}), is not in general conserved, though.
In $2+1$-dimensional Minkowski space, for example,
the $\xi$--preserving transformations do yield conserved
currents,
namely
the usual Schr\"odinger conserved quantities \cite{JHN},
\cite{barg}.
However, the currents associated to $\xi$--non--preserving
transformations
as antiboosts, etc. are manifestly not conserved, as seen from
(\ref{kchange}).
Let us conclude our investigations with explaining how the
results of
Jevicki \cite{JEV} fit into our framework.
Let us start with the free wave equation (\ref{waveq}) in $(2+1)$
dimensional Minkowski space
and let us assume that the scalar field has the form
\begin{equation}
\psi=\frac{1}{4\pi}\big[\Psi(x,t)e^{is}
+\Psi^\dagger(x,t)e^{-is}\big]
=
\frac{1}{2\pi}\sqrt{R(x,t)}\cos(\Theta+s),
\label{jevicki}
\end{equation}
where $\Psi(x,t)=\sqrt{R(x,t)}e^{i\Theta(x,t)}$.
This field is not equivariant but is rather a mixture of
two states with ``masses'' $(+1)$ and $(-1)$. Hence the usual
theory
of \cite{barg} does not apply. Nor does it fit perfectly
into our ``weaker'' theory~:
the phase is {\it identically} zero, so that any
$\Theta$ solves our equation (\ref{newequiv}).
Calculating the Lagrange density for
the Ansatz (\ref{jevicki}), we find, however,
\begin{equation}
\begin{array}{cc}
-2\pi{\cal L}_{0}
=
&\Big\{\2R\big({\partial}_{x}\Theta\big)^2+R{\partial}_{t}\Theta\Big\}
\sin^2(\Theta+s)
+
\displaystyle{\frac{({\partial}_{x}R)^2}{8R}
\cos^2(\Theta+s)}\hfill
\ccr
&-
\Big\{R{\partial}_{x}\Theta+{\partial}_{t}R\Big\}\sin(\Theta+s)
\cos(\Theta+s).
\hfill
\end{array}
\end{equation}
The vertical direction can be compactified with period $2\pi$. Then
integrating over $s$ yields the reduced action on ordinary
space--time
\begin{equation}
-\int dxdt
\left[\Big\{\2R\big({\partial}_{x}\Theta\big)^2
+R{\partial}_{t}\Theta\Big\}
+
\frac{({\partial}_{x}R)^2}{8R}\right].
\end{equation}
Removing the effective potential $\frac{({\partial}_{x}R)^2}{8R}$,
we end up
with the expression in (\ref{BJ}). It has therefore the same
${\rm O}(3,2)$ conformal symmetry.
\goodbreak
\parag\underbar{Acknowledgements}.
We are indebted to C.~Duval for sending us his
unpublished notes \cite{Duval} and for many enlightening
discussions.
We would like to thank also
D.~Bazeia, R.~Jackiw, A.~Jevicki, and N.~Mohameddi.
M.~H. acknowledges
the {\it Laboratoire de Math\'ema\-thi\-ques et de Physique
Th\'eorique}
of Tours University for hospitality, and
the French Government for a doctoral scholarship.
\goodbreak
|
\section{introduction}
The fractional quantum Hall (FQH) effect is a phenomenon observed in
a two-dimensional electron system subjected to a strong perpendicular
magnetic field.
Due to the interplay between the strong magnetic field and interactions
among the electrons as well as weak disorder, the transverse resistivity
shows a plateau behavior. \cite{tsui}
For a filling factor $\nu=1/$(odd integer), the theory predicts fractionally
charged quasiparticles with charge $q=\nu e$. \cite{bob}
Recent shot-noise experiments in a two-terminal FQH system with a
point-like constriction or a point quantum contact (QPC) between the edges
seem to be consistent with this theoretical
prediction.\cite{cglat}
The FQH system which have any experimental relevance should be confined
in a finite region enclosed by one or more edges.
Due to the presence of strong magnetic field, the
low-energy physics of this two-dimensional electron liquid reduces
essentially to that of the one-dimensional edge mode.
In one dimension it is known that the interaction plays
a significant role. The electrons are strongly renormalized
so that the Fermi liquid theory breaks down to be replaced
by the Tomonaga-Luttinger liquid (TLL).~\cite{hald}
If one considers spinless
electrons in 1D, the TLL is completely characterized by one
parameter $g$ which represents the strength of interaction.
Therefore the parameter $g$ for an interacting electron system
in 1D is not universal. On the other hand a remarkable feature
of FQH edge mode is that the parameter $g$ which controls
this 1D system is universal, since $g$ is related to the
topological nature of the bulk FQH liquid (FQHL).
For the edge mode of principal Laughlin states, the parameter
$g$ is simply given by the bulk filling factor $\nu$.
\cite{wen1}
The edge-tunneling experiment in FQH liquids has shown
that a chiral TLL is realized at the edge of FQHL. \cite{chang}
Indeed the chiral TLL theory has succeeded in the description of
non-linear $I-V$ characteristics for $\nu=1/$(odd integer),\cite{wen2}
but it is also true that the edge-tunneling experiment
cannot be explained by a naive TLL theory for other filling factors.
\cite{grayson,dhlee}
In particular, for the Jain's composite fermion hierarchy states
at filling factor
$\nu=m/(mp+\chi)$ ($m$: integer, $p$: even integer, $\chi=\pm 1$),
\cite{jain}
Wen's chiral TLL theory predicts that there should be $m$
edge modes corresponding to each composite fermion Landau level.
\cite{wen3}
Due to the existence of these internal degrees of freedom
the predicted exponent $\alpha$ for the $I-V$ characteristics
does not fit the experiment: $\alpha\sim 1/\nu$.
Although the observed exponent $\alpha\sim 1/\nu$ for the tunneling
into FQHL does not support the hierarchial structure of edge mode,
there is another experimental observation which encourages us to
work on this theory. It is the suppressed shot-noise
measurement at bulk filling factor $\nu=2/5$, i.e., at $m=2,p=2,\chi=1$
in a constricted two-terminal Hall bar geometry. \cite{rez}
They observed the transitions of two-terminal conductance
from a plateau at $G=2/5$ to another at $G=1/3$ and finally to $G=0$
as the constriction is increased.
On the plateau at $G=1/3$ they observed a fractional charge $q=e/3$,
which indicates that the filling factor near the quantum point
contact (QPC) is $\nu=1/3$.
The experiment clearly indicates a deep connection
between the $\nu=2/5$ daughter state and the $\nu=1/3$ parents state,
and therefore seems to support the hierachy theory at $\nu=2/5$.
This paper studies the tunneling through a QPC at the edge of
FQHL. This topic has captured a widespread attention both
experimentally and theoretically. For the reasons stated above
we forcus on the filling factor $\nu=m/(mp+\chi)$.
The FQH systems at those filling factors will provide
an interesting arena to study the hierarchical
nature of those liquids.
We discuss the successive transition between the plateaux
of conductance. The sequence begins with the conductance
$G=G_m=m/(mp+\chi)$ in units of $e^2/h$, which
is identical to the bulk filling factor. We discuss
the selection rule for the transitions between different $G$'s.
Even though the eventual correctness
of the hierachical TLL theory description is yet to be
tested, on which we will be based, we insist that
it is of importantance to make various interesting
applications of the theory. It will enable us to compare the
experiment with the theoretical predictions, and hence
will be useful to judge the correctness of hierarchical picture.
\section{model}
Our model has a two-terminal Hall bar geometry.
The bulk FQHL in the $xy$-plane is confined electro-statically
on the $y$-direction into a finite region: $-w/2<y<w/2$.
Each end of this strip is connected to a source (the left terminal)
or to a drain (the right terminal).
We assume that the bulk FQHL is incompressible at a filling
factor $\nu$. Therefore
the low-energy excitations are allowed only in the vicinity
of two boundaries, which constitute the edge modes.
The upper (lower) edge mode carries a current from the
left (right) to the right (left) terminal, and the total
current $I$ is defined as the difference of two.
Since there is no mechanism of relaxation in the TLL itself,
the chemical potential is uniform in the respective edge modes, i.e.,
the upper (lower) edge mode has a chemical potential
equal to that of the source (drain).
All the scatterings occur inside the terminals.
\cite{land}
In the absence of point-like constriction, the two-terminal
conductance $G=I/V$ is quantized at $G=\nu$ in units of $e^2/h$,
since the back-scattering between the two edge modes
which breaks the momentum conservation is allowed
nowhere through the edge,
where $V$ is defined as the source-drain voltage.
Now we go back to the bulk FQHL. We forcus on a filling factor
$\nu=m/(mp+\chi)$ in the Jain's composite fermion hierarchy series,
where $m$: integer, $p$: even integer and $\chi=\pm 1$.
According to the bulk hierarchy structure, there should be $m$
edge modes, i.e., each edge mode corresponds to a composite
fermion Landau level in the bulk. Then
the low-energy physics of this electron liquid is
controlled by the $m$-channel edge mode, which obey the following
Lagrangian density, \cite{wen3}
\begin{equation}
{\cal L}_{\rm TLL}=
{i\over 4\pi}K^{\alpha\beta}
{\partial\phi_\alpha^+\over\partial\tau}
{\partial\phi_\beta^-\over\partial x}
+{1\over8\pi}U^{\alpha\beta}
\left(
\frac{\partial\phi_\alpha^+}{\partial x}
\frac{\partial\phi_\beta^+}{\partial x}
+
\frac{\partial\phi_\alpha^-}{\partial x}
\frac{\partial\phi_\beta^-}{\partial x}
\right),
\end{equation}
where $\phi^\pm=\phi^u\pm\phi^l$ with $\phi^u (\phi^l)$ being the edge mode
propagating near the upper (lower) boundary of the system.
The matrix $K$ in Eq.~(1) could be identified as
the so-called $K$-matrix in the bulk, which together with
the electromagnetic-charge vector $t$ completely classify the universal
properties of bulk FQHL.
\cite{zee}
The standard construction for the $K$-matrix at a hierarchical
filling factor $\nu=m/(mp+\chi)$
yields
\begin{equation}
K=K(m,p,\chi)=\chi I_m +pC_m
\end{equation}
in the unitary basis $t^T=(1, \cdots, 1)$,
where $I_m,C_m$ are $m\times m$ identity and pseudo-identity matrices.
By a linear transformation one can decompose the modes into
charge and pseudo-spin bosons.
Each row and column of the matrices corresponds to a Landau level for
the composite fermions, i.e., $\alpha, \beta=1, \cdots, m$.
However it would be fair to comment that it is still a controversial
question what the correct construction of the $K$-matrix is.
\cite{lopez}
The matrix $U$ in Eq.~(1) is a positive definite matrix,
which specifies among others the velocities of the edge modes.
For $\chi=1$ charge and pseudo-spin modes propagate in the same
direction (co-propagate),
whereas for $\chi=-1$ they are counter-propagating, i.e.,
$\chi$ stands for the chirality of the edge modes.
For the latter case ($\chi=-1$), the interaction between
the edge modes can make the conductance non-universal.
The observed conductance, on the contrary, seems to be
universal. A remedy for this puzzle would be to put disorder
along the edge.~\cite{KFP} In the presence of such disorder
our conclusions will be modified, however, which will
not be discussed in the body of the paper.
Now we introduce the back-scattering by pinching the Hall bar,
i.e., by breaking the global translational invariance at $x=0$.
Let us think of applying a gate voltage locally in the middle
of Hall bar. It squeezes the Hall bar and makes a quantum
point contact (QPC) between the two edges.
The QPC introduces the tunneling of quasiparticle through the
pinched region of Hall bar.
In the TLL model it corresponds to a backward scattering
and hence could be described by a periodic
potential barrier for the bosonic fields.
\cite{KFFN}
Let us remember that we are focusing on the bulk filling factor
$\nu_{\rm bulk}=m/(mp+\chi)$.
According to the Jain's composite fermion hierarchy,
there should be $m$ filled composite fermion Landau levels
in the bulk, and accordingly $m$ types of elementary
quasiparticles. Each correspond to a vortex-charge vector
$l=l_j$ where $(l_j)^\alpha=\delta_j^\alpha$ ($j,\alpha =1,\cdots,m$)
with $\delta_j^\alpha$ being unity for $\alpha =j$ and vanishes
otherwise.~\cite{blok} The fractional charge carried by the
quasiparticle $l$ is given in general as
\begin{equation}
q/e=t^T K^{-1} l = {1\over mp+\chi}\sum_{\alpha=1}^m l^\alpha.
\label{charge}
\end{equation}
For the elementary quasiparticles $l=l_j$ one finds
$q=e/(mp+\chi)$, which is indeed the smallest possible value.
For $m=2,p=2,\chi=1$, i.e., $q=e/5$ they could be identified
as the current-carrying particles observed in the recent
shot-noise experiment at $\nu=2/5$. \cite{rez}
The tunneling of quasiparticle on the $j$-th composite fermion
Landau level induces a potential barrier proportional to
$\delta (x)\cos\phi_j^{+}$. Since the scattering amplitude
would be different for different types of quasiparticles,
the tunneling of these elementary quasiparticles sums up to
the following scattering potential barrier:
\begin{equation}
{\cal L}_{\rm tun}^{\rm (I)}=\sum_{j=1}^m
u_j\delta (x)\cos\phi_j^{+}.
\label{I}
\end{equation}
We call it the scattering potential due to the tunneling
of Class [I] quasiparticles.
In Eq. (\ref{I}) we took into account only the `intra-Landau-level'
processes.
Now I draw your attention to another class of quasiparticles,
which we call Class [II]. It consists of $m$ elementary quasiparticles.
The vortex-charge vector assigned to this Class [II] quasiparticle
is $l^T=(1,\cdots,1)$. In this combination of the bosonic fields
all the neutral modes cancel and the tunneling of such
quasiparticles do not accompany any neutral modes.
The fractional charge carried by this quasiparticle is
found to be $q/e=m/(mp+\chi)=\nu$.
In the bosonic language it can be written as
$\phi_c=\phi_1+\cdots+\phi_m$, which indeed corresponds to
the charge mode.
The scattering potential due to the Class [II] quasiparticle
tunneling operator can be written as
\begin{equation}
{\cal L}_{\rm tun}^{\rm (II)}=u\delta (x)\cos\phi_c^{+}.
\label{II}
\end{equation}
Now our total Lagrangian density reads
${\cal L}_{\rm total}={\cal L}_{\rm TLL}+{\cal L}_{\rm tun}^{\rm (I)}
+{\cal L}_{\rm tun}^{\rm (II)}$.
In the RG analysis in Sec. IV, we study the scaling behavior
of $u_j$'s and $u$, which are controlled by the scaling dimesions
of the quasiparticle tunneling operators: $\cos\phi_j^{+}$
and $\cos\phi_c^{+}$. They are given by
$\Delta_I=\nu/m^2+1-1/m$ for the Class [I] quasiparticles,
whereas $\Delta_{II}=\nu$ for the Class [II]
quasiparticle, where $\nu=m/(mp+\chi)$.~\cite{wen2}
If the scaling dimension is smaller than 1, the corresponding
tunneling amplitude tends to have stronger values as the
voltage or the temperature is lowered.
It is indeed the case both for $u_j$'s and $u$.
In the parameter region $\{(m,p)|m\geq 2, p\geq 2\}$ in which we are
interested, one can prove that
\begin{equation}
\left\{
\begin{array}{l}
\Delta_I=\Delta_{II}\ \ \ {\rm for}\ \ \ \chi=-1, m=2, p=2\ \ \ (\nu=2/3)
\\
\Delta_I>\Delta_{II}\ \ \ {\rm otherwise}
\end{array}
\right.,
\end{equation}
i.e., the Class [II] quasiparticles have a lower scaling dimension
in most of the cases, and hence more relevant in the RG sense.
Another important observation is that there would be at least
two ways how the scattering becomes stronger. One way is, as we
have described above, to increase its amplitude. However it would
be also possible that higher-order cascade of scatterings becomes
important, where the single QPC description is no longer valid.
One might have to take into account the resonance in such
non-perturbative regime.
\section{Hypotheses}
In Sec. II we gave expressions to the possible tunneling
processes through a single QPC in terms of the bosonic
field $\phi_j$ or of its linear combination $\phi_c$.
We considered two classes of quasiparticles. The Class [I]
corresponds to the elementary quasiparticles with the
smallest fractional charge.
The Class [II] corresponds to the charge mode:
$\phi_c=\phi_1+\cdots+\phi_m$.
We also compared the scaling dimensions of the two classes
of quasiparticle tunneling operators. Although the Class [II]
quasiparticles have a lower scaling dimension in most of
the cases and more relevant in the RG sense,
it is not unlikely that the Class [II]
is negligible for some reason; since they are
bound states of $m$ elementary quasiparticles,
they are so scarcely created that
the scattering potential (\ref{II}) could not develop enough to
be effective at the energy scales in question despite its
relevant scaling dimension.
Therefore we are encouraged to consider the following cases:
\begin{enumerate}
\item
Case [A]:
Class [II] quasiparticles are negligible for some reason.
Furthermore the tunneling amplitudes for
each channel $j$ have different orders of magnitude:
\begin{equation}
u_m\gg u_{m-1}\gg\cdots\gg u_1.
\label{A}
\end{equation}
\item
Case [B]:
Class [II] is still negligible, but some $u_j$'s have comparable
orders of magnitude;
\begin{equation}
u_{j+1}\gg u_j\sim\cdots\sim u_{j-k+1}\gg u_{j-k},
\label{B}
\end{equation}
where $k\geq 2$ is an integer.
\item
Case [C]:
Class [II] is no longer negligible, i.e.,
the amplitudes $u$ for the Class [II] quasiparticle has a comparable
magnitude with those for $u_j$'s.
\end{enumerate}
The assumption (\ref{A}) for Case [A] might be justified in a way analogous
to the edge-channel argument for integer quantum Hall effect
(IQH). \cite{halp}
Let us consider the Landau levels for composite fermions.
The lowest $m$ Landau levels are completely filled by the
composite fermions, and the chemical potential lies
between the $m$-th and $(m+1)$-th Landau levels.
Towards the edge of the sample each energy level tends to be
lifted up by the confining potential.
Since the $j$-th edge mode lies in where the chemical potential crosses
the $j$-th energy level, each edge channels are spatially
separated. Therefore the tunneling between the
$m$-th edge modes, the spatially closest ones, is supposed to have
a much larger amplitude than the other $m-1$ channels.
It is also the case for $u_{m-1}$ compared with the remaining
$m-2$ channels and so forth.
I would like to mention an experiment which encourages us to
employ the assumptions (6).
It is the suppressed shot-noise
measurement at bulk filling factor $\nu=2/5$, i.e., at $m=2,p=2,\chi=1$.
~\cite{rez}
They observed the transitions of two-terminal conductance
from a plateau at $G=2/5$ to another at $G=1/3$ and finally to $G=0$
as the constriction is increased.
On the plateau at $G=1/3$ they observed a fractional charge $q=e/3$,
which indicates that the filling factor near the QPC is $\nu=1/3$.
Hence a single-channel edge mode described by a
$1\times 1$ $K$-matrix; $K=3$ is expected near the QPC.
~\cite{nom}
On the other hand in the region where the physics is completely
unaffected by the gate, the matrix $K$ should be given by
$K=K(2,2,1)$.
This experiment not only indicates a deep connection
between the $\nu=2/5$ daughter state and the $\nu=1/3$ parents state.
It also implies that $u_2\gg u_1$ as well as $u$ is negligiblly small.
We have explained above a physical reason why we are intereted
in the parameter region (\ref{A}). However we have another example
where the assumption (\ref{B}) seems to be reasonable.
It is the spin-singlet state at $\nu=2/3$
($m=2,p=2,\chi=-1$), where $j=1,2$ corresponds to each spin indices
instead of each Landau level for the spin-polarized state.
If we neglect the Zeeman energy, the theory should be symmetric
in terms of the two spin components.
Therefore it seems more reasonable to assume in this case as
$u_1\sim u_2$, which belongs to our Case [B].~\cite{rapid}
\section{RG flow and critical phenomena}
Let us forget for the moment the Class (II) quasiparticles.
Then we consider the renormalization group (RG) phase diagram in the
$m$-dimensional space of $u =(u_1, \cdots, u_m)$.
The origin of this plane
corresponds to the conductance plateau at
$G=m/(mp+\chi)=\nu_{\rm bulk}$.
A standard RG analysis shows that
only the fixed point at $u=(\infty,\cdots,\infty)$,
is infra-red (IR) stable,
since all $u_j$'s are found to be relevant.
We introduce a small negative gate voltage to the system
on the conductance plateau at $G=G_m$.
We fix the gate voltage
so that the scaling at the zero temperature
should be controlled by the voltage difference $qV$
between the two reservoirs. Now we ask where the initial
point of our RG flow is.
\subsection{Successive transistions}
Let us consider the Case [A]. Our RG flow starts from a point
in the vicinity of the origin (unstable fixed point)
where Eq. (\ref{A}): $u_m\gg u_{m-1}\gg\cdots\gg u_1$ is satisfied.
As the voltage is decreased, all $u_j$'s scale to
larger values. But due to the assumption (\ref{A})
$u_m$ increases much faster than the other $m-1$ channels.
Therefore our RG path flows into the domain $D_{m-1}$, where
$D_j$ ($j=1,\cdots,m-1$) is defined as
\begin{equation}
D_{j}=\{(u_1,\cdots,u_m)|
u_m,\cdots,u_{j+1}>u_{\rm crtc}\gg u_{j},\cdots,u_1\}.
\end{equation}
$u_{\rm crtc}$ is a critical value of the tunneling
amplitudes such that the phase $\phi_j$ is pinned
when $u_j>u_{\rm crtc}$
in order for $g$ to be quantized.
In reality $u_{crtc}$ is determined by the strength of impurity
potential which could retain the induced quasiparticles
at the impurity cite.
In the domain $D_{m-1}$ the effective $K$-matrix near the PC
reduces to $K=K(m-1,p,\chi)$.
Therefore one could indentify $D_{m-1}$
to the plateau of conductance at $G=G_{m-1}$.
However since the domain $D_{m-1}$ corresponds
to a saddle region of the RG flow,
our RG path flows away from $D_{m-1}$
and goes toward the next saddle region $D_{m-2}$
defined in the same way as $D_{m-1}$.
We further introduce the domain $D_j$ in general for
$j=1,\cdots,m-1$.
Our RG flow passes through $D_j$'s as
\begin{equation}
D_{m-1}\rightarrow D_{m-2}\rightarrow\cdots\rightarrow D_1,
\end{equation}
and finally it flows into the attractive
fixed point $u=(\infty,\cdots,\infty)$,
which is identified as the completely reflecting phase $G=0$:
the domain $D_0$. (Fig. 1)
The exceptions are the series belonging to $\chi=-1, p=2$,
i.e., the $\nu=2/3$ state and its daghter states.
For those filling factors our RG stops at the
$G=1$ plateau.~\cite{nom,PRB}
In the following we consider the other cases.
As the RG path flows from the vicinity of the origin toward
the $G=0$ phase, the effective $K$-matrix near the PC
changes as
\begin{equation}
K(m,p,\chi)\rightarrow K(m-1,p,\chi)\rightarrow \cdots \rightarrow
K(1,p,\chi)=p+\chi \rightarrow {\rm insulator}.
\end{equation}
Correspondingly we predict the following successive
plateau transitions,
\begin{equation}
G_m\rightarrow G_{m-1}\rightarrow\cdots\rightarrow G_1={1\over p+\chi}
\rightarrow 0.
\label{cond}
\end{equation}
One might think the above result is very close to
the `global phase diagram' in the quantum Hall effect.~\cite{KLZ}
Though it indeed is, it differs in that
the direction of the transition is specified in
our case.
Anomalous transitions ($G_j\rightarrow G_{j-k}$ for $k\geq 2$)
are forbidden as far as the assumption (6) is satified.
I would like to deduce the scaling behavior of the
tunneling current and the shot-noise spectrum on the plateau $G=G_j$
($j=1,\cdots,m$).
The back-scattering current
$I_b=\nu (e^2/h)V-I$
can be calculated perturbatively with respect to $u_{j-1}$
and obtained as~\cite{wen1}
\begin{equation}
\langle I_b \rangle
={2\pi q\over \Gamma[2\Delta_I]}|u_{j-1}|^2
{a^{2\Delta_I-2}\over v_c^{2\nu/m^2}v_s^{2(1-1/m)}} (qV)^{2\Delta_I -1},
\end{equation}
where $q=e/(jp+\chi)$ is a fractional charge of the
elementary quasiparticle on the plateau $G=G_j$, and
$\Delta_I$ is a scaling dimension of the Type (I) quasiparticle tunneling
operator: $\Delta_I=\nu/m^2+1-1/m$.
$a$ is a short-distance cutoff, and $v_c$ and $v_s$ are velocities
of the charge and the pseudo-spin modes, respectively.
The shot-noise spectrum
\begin{equation}
S (\omega)=\int_{-\infty}^{\infty}dt\cos\omega t
\langle\{I_b(t),I_b(0)\}\rangle
\end{equation}
is also calculated perturbatively to give,
$S (\omega)=q \langle I_b \rangle
(|1-{\omega/qV}|^{2\Delta_I-1}+
|1+{\omega/qV}|^{2\Delta_I-1})$,
which reduces to $S=2q \langle I_b \rangle$
in the white-noise limit ($|\omega|\ll qV$).~\cite{chmn}
They are also calculated near the insulating phase to be
$S(\omega)=2e\langle I\rangle$ with
\begin{equation}
\langle I \rangle
={2\pi e\over \Gamma[2(p+\chi)]}|\tilde{u_1}|^2
a^{2(p+\chi)-2} \tilde{v_1}^{-2(p+\chi)} (eV)^{2(p+\chi)-1},
\end{equation}
where $\tilde{u_1}$ represents
the strength of electron tunneling dual to $u_1$
and $\tilde{v_1}$ a corresponding velocity.
\subsection{Anomalous transition}
Let us turn to the case [B], where the Type [II] quasiparticles are
still negligible, but some $u_j$'s have comparable orders of
magnitude.
Here we consider a particular case of the assumption (7);
we consider the case where all $u_j$'s have comparable orders of
magnitude:
\begin{equation}
u_m\sim u_{m-1}\sim\cdots\sim u_1.
\label{B2}
\end{equation}
In this case a direct transition from $G=G_m$ to $G=0$
is expected, since all $\phi_j$'s tend to be pinned
at the same speed. The shot-noise spectrum on the plateau
at $G=G_m$ is given by $S(\omega)=2q\langle I_b\rangle$
for $|\omega|\ll qV$ with
\begin{equation}
\langle I_b \rangle
={2\pi q\over \Gamma[2\Delta_I]}
\sum_{j=1}^{m}|u_{j}|^2
{a^{2\Delta_I-2}\over v_c^{2\nu/m^2}v_s^{2(1-1/m)}} (qV)^{2\Delta_I -1}.
\end{equation}
Now we turn our discussion to the insulating phase: $G=0$.
Remember each vortex-charge vector $l$ with integer elements
corresponds to a quasiparcile which has a charge given by
(\ref{charge}).
To construct an electron operator, we have only to set
Eq. (\ref{charge}) to be equal to 1. Of course, there is
in priciple an infinite number of choice of $l$ to
make it identical to unity.
However, as far as the tunneling is concerned,
we can pick up most relevant electron operators,
which are found to be \cite{blok}
\begin{equation}
l=\tilde{l}_j,\ \ \ (\tilde{l}_j)^\alpha=p+\chi\delta_j^\alpha,
\end{equation}
where $j=1,\cdots,m$ and each $\tilde{l}_j$
has $m$ components, i.e., $\alpha=1,\cdots,m$.
These electrons look analogous to our Class [I] quasiparticles.
It indeed is, but we will see that the relation is deeper.
Before going into that, the scattering potential barrier
due to the tunneling of these `Class [I] electrons'
can be written as
\begin{equation}
\tilde{{\cal L}}_{\rm tun}^{\rm (I)}=\sum_{j=1}^m
\tilde{u}_j\delta (x)
\cos\left[
\sum_{\alpha=1}^m (\tilde{l}_j)^\alpha \tilde{\phi}_\alpha^{+}
\right].
\label{ET}
\end{equation}
Here the `inter-Landau-level' tunnelings are neglected again.
Note that the $x$-axis is taken along the edge which is
assumed to be completely reflected in the insulating phase.
Let us take notice of the duality between
the quasiparticle tunneling and the electron tunneling,
which is exact when $u_m =u_{m-1}=\cdots =u_1$.
~\cite{rapid}
To see this, let us go back to the weak-scattering phase,
i.e., we start with the Lagrangian density:
${\cal L}_{\rm total}={\cal L}_{\rm TLL}+{\cal L}_{\rm tun}^{\rm (I)}$.
We starts our RG from the vicinity of the origin
in the $m$-dimensional space of $u =(u_1, \cdots, u_m)$.
We assume that the condition (\ref{B2}) is satisfied.
As the voltage is decreased, all $u_j$'s scale to
larger values. Then one is encouraged to employ the duality
transformation, i.e., one considers the tunneling of
instantons between the potential minima.~\cite{schmid}
Up to the lowest non-trivial order with respect to
those instantons, one obtains a model which has exactly
the same form as Eq. (\ref{ET}),
where $-\tilde{u}_j /2$ corresponds to an instanton fugacity
whereas \{$\tilde{\phi}_\alpha^{+}$\} is identified as a set of
bosonic fields dual to \{$\phi_\alpha^{+}$\}.
The shot-noise spectrum in the insulating phase
is given by $S(\omega)=2e\langle I\rangle$ for $|\omega|\ll qV$
with the tunneling current $I$ scaling as
\begin{equation}
\langle I \rangle
={2\pi e\over \Gamma[2\tilde{\Delta}_I]}
\sum_{j=1}^{m}|\tilde{u}_j|^2
{a^{2\tilde{\Delta}_I-2}\over
\tilde{v}_c^{2/\nu}\tilde{v}_s^{2(1-1/m)}}
(eV)^{2\tilde{\Delta}_I-1}.
\end{equation}
The scaling dimension $\tilde{\Delta}_I$ of our Class [I] electron
tunneling operator is given by $\tilde{\Delta}_I=1/\nu+1-1/m$.
$\tilde{v}_c,\tilde{v}_s$ are velocities in the insulating phase.
\subsection{Direct transition}
Let us consider the case [C]. In this case we obtain
still different results.
In the presence of Class [II] quasiparticles,
the physics tends to be controlled by the scattering potential
$u$ as the energy in question is lowered.
In the region where $u_1,\cdots,u_m\ll u\ll qV$ is
satisfied, i.e., $G\sim G_m$, one obtains
$S(\omega)=2q\langle I_b\rangle$ with
\begin{equation}
\langle I_b \rangle
={2\pi q\over \Gamma[2\Delta_{II}]}|u|^2
a^{2\Delta_{II}-2} v_c^{-2\Delta_{II}} (qV)^{2\Delta_{II}-1}.
\end{equation}
The fractional charge of the Class [II] quasiparticle
is identical to the bulk filling factor:
$q/e=m/(mp+\chi)=\nu$, which is also equal to the scaling
dimension of the corresponding quasiparticle tunneling operator:
$\Delta_{II}=m/(mp+\chi)=\nu$.
As the RG path flows into the strong-scattering phase,
the conductance $G$ shows a direct transition to $G=0$ again.
However the scaling behavior of the tunneling current
in the insulating phase is less clear. The reason is that
the duality is not existing in this case so that the
physical interpretation of the strong-scattering phase Lagrangian
is lacking.
\section{summary: the selection rule}
In summary we obtained the following selection rules
for the transition between plateaux starting with
the bulk value $G=m/(mp+\chi)=\nu$.
For Case [C] a direct transition to the Hall insulator
is expected.
For Cases [A] (and [B]) successive transitions from one $G$
to another $G$ is allowed under the following
selection rule:
\begin{eqnarray}
g=\nu(j,p,\chi)&\rightarrow&g=\nu(j',p',\chi')
\nonumber \\
j'=j-1,\ \ p'&=&p,\ \ \chi'=\chi.
\end{eqnarray}
An anomalous transition $j\rightarrow j-k$ ($k>2$: integer) is expected
between the same $p$ and $\chi$ when the condition (\ref{B})
is satisfied.
The overall picture of the system which
results from the plateaux transitions discussed above is the following.
We started with the situation where the filling factor
is extended uniformly over the whole system, i.e., equal to the bulk
value $\nu=m/(mp+\chi)$.
Then we effectively increased the gate voltage
in units of the voltage difference $V$ between the two terminals.
We found successive transitions of the conductance
(\ref{cond}) when the condition (\ref{A}) is satisfied.
The question is what happens
between the QPC and the bulk FQHL.
Each time $G$ passes through one plateau
($G=G_j$), there should appear one additional incompressible strip
with a filling factor $\nu=j/(jp+\chi)$.
Before ending this paper I mention that the edge-confining potential
is assumed to be steep through the paper enough to avoid the complexities
which may arise when the confining potential is smooth.
\cite{been,chkl}
In conclusion we studied the successive transitions of conductance
between different plateaux of hierarchical FQHL.
The scaling behavior of the tunneling current
and the shot-noise spectrum are calculated perturbatively
on each plateau of the conductance.
We discussed the selection rules
for the transition between different plateaux of the conductance
in order that the theory could be tested by the experiments.
\acknowledgements
This paper is an extension and a generalization of an earlier
paper with K. Nomura (Ref.~\cite{nom}).
I am grateful to him for useful discussions and a collaboration.
I am also grateful to Y. Morita for a key comment to
initiate the present work.
I would like to thank P. Lederer and N. Nagaosa for their
suggestions and encouragements.
I would like to acknowledge the kind hospitality
during my participance in the Trimestre Fermions Fortement
Correles (IHP, Paris).
I was supported by JSPS Research Fellowships for Young Scientists,
and am supported by Ministere de l'Education Nationale,
de la Recherche et de la Technologie.
\references
\bibitem{tsui}
D.C. Tsui, H.L. Stoemer and A.C. Gossard,
Phys. Rev. Lett. {\bf 48} 1559 (1982).
\bibitem{bob}
R.B. Laughlin,
Phys. Rev. Lett. {\bf 50} 1395 (1982).
\bibitem{cglat}
L. Saminadayar, D.C. Glattli, Y. Jin and B. Etienne,
Phys. Rev. Lett. {\bf 79} 2526 (1997);
R. de-Picciotto, M. Reznikov, M. Heiblum,V. Umanski,
G. Bunin and D. Mahalu, Nature {\bf 389}, 162 (1997).
\bibitem{hald}
F.D.M. Haldane, Phys. Rev. Lett. {\bf 45}, 1358 (1980);
ibid. {\bf 47}, 1840 (1981).
\bibitem{wen1}
X.-G. Wen, Phys. Rev. {\bf B41}, 12838 (1990).
\bibitem{chang}
A. M. Chang, L. N. Pfeiffer, and K. W. West,
Phys. Rev. Lett. {\bf 77}, 2538 (1996).
\bibitem{wen2}
X.-G. Wen, Phys. Rev. {\bf B44}, 5708 (1991);
K. Moon, H. Yi, C.L. Kane, S.M. Girvin, and M.P.A. Fisher,
Phys. Rev. Lett. {\bf 71}, 4381 (1993);
C. de Chamon and E. Fradkin, Phys. Rev. {\bf 56}, 2012 (1997).
\bibitem{grayson}
M. Grayson, D.C. Tsui, L.N. Pfeiffer, and K.W. West,
and A. M. Chang, Phys. Rev. Lett. {\bf 80}, 1062 (1998).
\bibitem{dhlee}
D.H. Lee and X.-G. Wen, cond-mat/9809160;
K.-I. Imura, Europhys. Lett. {\bf 47}, 233 (1999).
\bibitem{jain}
J.K. Jain, Phys. Rev. Lett. {\bf 63}, 199 (1989).
\bibitem{wen3}
X.-G. Wen, Adv. Phys. {\bf 44}, 405 (1995).
\bibitem{rez}
M. Reznikov, R. de-Picciotto, T.G. Griffiths, M. Heiblum and V. Umanski,
Nature {\bf 399}, 238 (1999).
\bibitem{land}
R. Landauer, Phil. Mag. {\bf 21}, 863 (1970);
M. Buettiker, Phys. Rev. {\bf B38} 9375 (1988).
\bibitem{zee}
J. Froehlich and A. Zee,
Nucl. Phys. {\bf B364}, 517 (1991);
X.G. Wen and A. Zee,
Phys. Rev. {\bf B46}, 2290 (1992).
\bibitem{lopez}
A. Lopez and E. Fradkin, cond-mat/9810168.
\bibitem{KFP}
C.L. Kane, M.P.A. Fisher and J. Polchinski Phys. Rev. Lett.
{\bf 72}, 4129 (1994).
\bibitem{KFFN}
C.L. Kane and M.P.A. Fisher, Phys. Rev. {\bf B46}, 15233 (1992);
A. Furusaki and N. Nagaosa, Phys. Rev. {\bf B47}, 3827 (1993).
\bibitem{blok}
B. Blok and X.-G. Wen, Phys. Rev. {\bf B42}, 8133 (1990).
\bibitem{halp}
B.I. Halperin, Phys. Rev. {\bf B25}, 2185 (1982).
\bibitem{nom}
K.-I. Imura and K. Nomura, Europhys. Lett. {\bf 47}, 83 (1999).
\bibitem{rapid}
K.-I. Imura and N. Nagaosa, Phys. Rev. {\bf B57}, R6826 (1998).
\bibitem{PRB}
K.-I. Imura and N. Nagaosa, Phys. Rev. {\bf B55}, 7690 (1997).
\bibitem{KLZ}
S. Kivelson, D.-H. Lee and S.-C. Zhang, Phys. Rev. {\bf B46}, 2223 (1992).
\bibitem{chmn}
C.L. Kane and M.P.A. Fisher, Phys. Rev. Lett. {\bf 72}, 724 (1994);
C. de Chamon, C. Freed and X.G. Wen,
Phys. Rev. {\bf B51}, 2363 (1995);
P. Fendley, A.W.W. Ludwig and H. Saleur,
Phys. Rev. Lett, {\bf 75} 2196 (1995).
\bibitem{schmid}
A. Schmid, Phys. Rev. Lett. {\bf 51}, 1506 (1983).
\bibitem{been}
C.W.J. Beenakker, Phys. Rev. Lett. {\bf 64}, 216 (1990);
A.M. Chang, Solid State Commun. {\bf 74}, 871 (1990).
\bibitem{chkl}
Y. Meir, Phys. Rev. Lett. {\bf 72}, 2624 (1994);
L. Brey, Phys. Rev. {\bf B 50}, 11861 (1994);
D.B. Chklovskii, Phys. Rev. {\bf B 51}, 9895 (1995).
\begin{figure}[h]
\input epsf
\epsfxsize=15.0cm
\epsfbox{rg.eps}
\vspace{1.0cm}
\caption{RG phase diagram in the $u$-space for $m=3$,
$u=(u_1,u_2,u_3)$:
Case 1 represents a RG flow corresponding to the
successive transitions which are expected when
$u_3\gg u_2\gg u_1$.
Case 2 represents a direct transition to the completely reflecting
phase $D_0$, which happens when
$u_3\sim u_2\sim u_1$.
A direct transition to $D_0$ is also expected
for Case 3, where the Type (II) scattering potential plays
the role.}
\end{figure}
\end{document}
\end
|
\section{Introduction}
The gravitational field of the inhomogeneous distribution of mass in
the Universe produces observable deviations from the smooth Hubble
expansion. Well-determined distances to galaxies provide an
opportunity to measure their motions relative to the ``Hubble flow''
-- so-called peculiar velocities -- which can lead to mass estimates
for a variety of systems, as in studies of the Local Group and of the
Virgocentric infall, or on larger scales via comparisons of peculiar
velocity measurements to expectations from full-sky redshift surveys (Dekel
1994, Willick \& Strauss 1995).
In such analyses, the Centaurus region is probably the most perplexing
zone of large-scale flow in our vicinity. It has a complex spatial
structure, and its peculiar velocity has been measured in some studies
to be much higher than that expected from the observed galaxy density.
Lucey, Currie, and Dickens (1986b) first called attention to the
apparently bimodal nature of the Centaurus cluster at ({\it l,b}) =
(302$^\circ$, 22$^\circ$), dividing it into two pieces at apparent
redshifts in the Local Group (LG) reference frame of approximately 2800 and
4300 $\,{\rm km\,s^{\scriptscriptstyle -1}}$\ (Cen30 and Cen45, respectively). In a deeper study of the
central portions of the cluster, Stein et al. (1997) found that dwarf
galaxies in Centaurus exhibit a clear concentration around the
redshift of NGC~4696 (an elliptical galaxy which is the brightest in
the cluster), $v_{lg}=2674 \pm 26$ $\,{\rm km\,s^{\scriptscriptstyle -1}}$, tracing a galaxy cluster
of velocity dispersion $933 \pm 118$ $\,{\rm km\,s^{\scriptscriptstyle -1}}$\ that they identify with
Cen30. Cen45, they determined, more strongly resembles a group falling
into Cen30, with a small velocity dispersion ($131 \pm 43$ $\,{\rm km\,s^{\scriptscriptstyle -1}}$) and a
population dominated by late-type galaxies.
A number of secondary distance
indicators have by now been applied to Centaurus galaxies, with often
contradictory results. Aaronson {\it et al.\/} (1989) were the first
to obtain distances for Centaurus spiral galaxies; they measured
peculiar velocities of $-80\pm 250$ and $+10\pm 450$ in the Local
Group reference frame for Cen30 and Cen45, respectively.
However, obtaining Tully-Fisher distances to these clusters is somewhat
problematic. Because of the large velocity dispersion of Cen30 and
the relatively small number of galaxies in Cen45, separating cluster
from background spirals is very difficult (see Lucey, Currie and
Dickens 1986a and Giovanelli et al. 1997 for examples). Reflecting
these difficulties, Aaronson et al. identify 6 galaxies spanning
nearly 2 magnitudes in distance modulus as belonging to Cen45. In contrast, based on $D_n - \sigma$ observations of elliptical galaxies, Faber et
al. (1989) reported the peculiar velocities for Cen30 and Cen45 to be
$+527\pm 214$ $\,{\rm km\,s^{\scriptscriptstyle -1}}$\ and $+1090\pm 336$ $\,{\rm km\,s^{\scriptscriptstyle -1}}$\ in the Local Group frame.
In an attempt to resolve such contradictory estimates of the distance
to the Centaurus region, we have undertaken a search for Cepheids in
the spiral galaxy NGC~4603 to firmly establish its location. This
galaxy is located near the center of the Cen30 cluster in position on
the sky, and has a velocity $v_{lg}=2321 \pm 20$ $\,{\rm km\,s^{\scriptscriptstyle -1}}$\ (Willick et
al.), well within the velocity dispersion of the cluster. NGC~4603
has an inclination of 53$^\circ$ and a 21cm width (20\%) of 411 $\,{\rm km\,s^{\scriptscriptstyle -1}}$\
with isophotal ($D_{25}$) diameter of 1.6'; Aaronson {\it et al.\/}
(1989) show that it fits onto their IRTF relation for Cen30 galaxies
quite well, with a distance modulus within 0.07 magnitudes
(0.3$\sigma$) of that derived for the cluster. It thus seems an
appropriate choice for such a study. Such a study should also allow
tests of the validity of the uniformity of the Tully-Fisher or $D_n -
\sigma$ relationships to a greater distance than has been possible
before.
However, even the smaller estimates of the distance to Cen30 place it
substantially further than any galaxy for which a search for Cepheids
has been previouly attempted, even using the Hubble Space Telescope;
the greatest distance modulus previously measured with this method is
that to NGC~4639, $32.03\pm 0.22$ ($25.5\pm2.5$ Mpc; Saha et
al. 1997). The redshift of Cen30 is roughly twice that of the Virgo
or Fornax clusters. It is thus reasonable to expect that observing
Cepheids in NGC~4603 should be difficult; not only do more distant
Cepheids appear fainter, but also the crowding of stars that
complicates photometry becomes more severe as the angular size
distance increases. Furthermore, the Centaurus cluster lies behind a
zone of substantial ($A_V\sim 0.5$) Galactic extinction, making any
stars observed that much fainter. In this regime, photometric errors
are significant enough that nonvariable stars have an appreciable
probability of appearing to vary in a manner indistinguishable from
a Cepheid with significant amplitude. Such obstacles might be overcome
by observing at many more epochs or with a greater exposure time per
epoch than in prior Cepheid studies, but the limited availability of
HST makes that infeasible.
Therefore, we have developed new techniques for dealing with such a
dataset. Instead of relying on a set of variability criteria for
preselection, we attempt to fit template Cepheid light curves to all
stars with well-determined photometry and then apply a series of
criteria that are effective at eliminating nonvariables. Even that
technique leaves a substantially contaminated list of candidate
Cepheids. We therefore do not obtain distance moduli from a direct
Period-Luminosity relation fit, but rather have developed a Maximum
Likelihood formulation that accounts for the properties of
nonvariables that mimic Cepheids and of the probability of selecting
an actual Cepheid of given properties based upon the results of
realistic simulations. These techniques allow us to minimize the
biases in distance determination that might otherwise appear and which
may have affected other Cepheid studies that have
pushed the limits of the technique.
We describe the details of the observations in $\S$ 2, the
procedures used to analyze the data and find Cepheids and our
simulations thereof in $\S$ 3, and our Maximum Likelihood formalism
and the determination of the distance to NGC~4603 in $\S$ 4.
\section{Observations}
We have observed NGC~4603 using the Wide Field and Planetary
Camera 2 (WFPC2) instrument on the Hubble Space Telescope (HST).
HST made a total of 11 distinct visits to the targeted field: 9,
totaling 24 orbits, using the F555W filter (roughly equivalent to
Johnson $V$), and 2, totaling 6 orbits, using the F814W filter
(similar to Kron-Cousins $I$). To ensure ease of data analysis, the
same orientation was maintained for all observations; the
telescope was generally dithered by 5.5 planetary camera pixels
($\approx$ 0\Sec25) between orbits. Two successive frames of
data were obtained during each orbit to minimize the effects of
cosmic rays. Due to technical limitations (such as the time
required to acquire the target field and the limited visibility of
NGC~4603 during the course of an orbit), the total integration
time was 900-1300 seconds per frame.
Our observing strategy was in general similar to that used for the
$H_0$ Key project (see, e.g., Freedman et al. 1994); however, due
to the large predicted distance of NGC~4603 ($>$20 Mpc), we could
expect to find only the longest period Cepheids (i.e., $P \lower.5ex\hbox{$\; \buildrel > \over \sim \;$}
25$ days). In fact, if NGC~4603 were located at $\lower.5ex\hbox{$\; \buildrel > \over \sim \;$}$45 Mpc,
Cepheids in this galaxy would be too faint to discover at all with
the WFPC2 instrument. We thus tried to optimize our observing
sequence to facilitate the discovery of longer-period variables.
Our original plan was to perform 8 F555W visits over the course of
$\approx 60$ days in 1996, spaced to maximize our ability to
detect and parameterize Cepheids with a variety of periods (as
described in Freedman et al. 1994). Unfortunately, the final
observation planned for 1996 did not occur due to an HST safing
event. Our sensitivity for the longest-period Cepheids --- exactly
those which are brightest and easiest to find --- is therefore
limited; those detected suffer from substantial aliasing in period
determination. Details of the observations performed are listed
in Table 1. In Figure 1, we show the results of a simulation for
the expected error in period determination as a function of period
for the sampling ultimately used, illustrating the effects of
aliasing.
\section{Data Analysis}
\subsection{Photometry}
The data were calibrated via the standard Space Telescope Science
Institute pipeline processing (Holtzman et al. 1995), applying the
Hill et al. (1998) long-exposure magnitude zero point. Each frame
was also corrected for vignetting and geometrical effects on the
effective pixel area as described in Stetson et al. (1998).\\
\vbox{%
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=8.9cm
\epsffile{f1.eps}}
\begin{small}
\figcaption{\small The percentage difference between the period measured using
our algorithms and the actual period for simulated observations of
stars with Cepheid light curves of a variety of phases and amplitudes
measured (with realistic magnitude errors appropriate for stars with
$F555W$=27.3-27.5) at the epochs of our actual observations, as a
function of period. The period errors are dominated by mistakenly
identifying an alias as the actual period. The diagonal striping
apparent here is due to the gridding in period used for fitting
template light curves.}
\end{small}
\end{center}}
Background levels in the data frames were high enough and exposure
times long enough that neither charge transfer inefficiencies nor
variations in the photometric zero points with exposure time
should significantly affect our results (cf. Rawson et al. 1997
and references therein). Each of the
WFPC2 chips was analyzed separately. Because the second WF chip
contained the nucleus of NGC~4603, crowding was severe and few
stars could be resolved in it; that chip was therefore omitted
from analysis. The fourth WF chip was directed at an outer
portion of the galaxy, containing few stars and no significant
numbers of Cepheids; it, too, was therefore removed from our
analysis.
Photometry was then performed on each of the data frames using the
DAOPHOT II/ ALLFRAME package (Stetson 1987). As an independent check,
magnitudes were also obtained using a version of DoPHOT (Schechter,
Mateo \& Saha 1993) modified by Abi Saha for use with HST data (see,
e.g., Ferrarese et al. 1996). The DoPHOT reductions were used as a
consistency check only; the analysis presented in this paper is based
on the ALLFRAME photometry alone. For F555W observations, the two
sets of photometry agreed to within $\pm$ 0.08 magnitudes on average;
this agreement is consistent with that found for distant galaxies
observed as part of the Key Project (e.g., Ferrarese et al. 1996,
Silbermann et al. 1998). The ALLFRAME analysis was more extensive
and resulted in larger numbers of Cepheid candidates; for candidates
found using both packages, the agreement in period was found to be
well within the errors quoted below.
For the ALLFRAME photometry, procedures similar to those of the HST
Key Project on the Extragalactic Distance scale were used (see, e.g.,
Kelson et al. 1996 for a more detailed description). ALLFRAME
performs photometry by fitting a predefined point-spread function
(PSF) to all stars on a frame and iteratively determining their
magnitudes. Files describing the WFPC2 PSF and its variation across
the field (determined from observations of globular clusters; cf. Hill
et al. 1998) were provided by P. Stetson. For each epoch, up to six
HST frames were obtained, and thus up to six measurements of each
star's magnitude were made. Those measurements are sometimes
contaminated by cosmic rays or other transient phenomena. Although
ALLFRAME attempts to limit their effect, it was found that simply
averaging the magnitudes determined using ALLFRAME and weighting them
according to their error estimates sometimes yields very inaccurate
results. We therefore experimented with a number of robust estimators
for the mean of the ALLFRAME measurements (including median, Tukey
biweight, and trimean; cf. Beers et al. 1990) using the magnitudes of
artificial stars inserted (using the ALLFRAME PSF) on our data
frames. The most successful proved to be an iterative reweighting
method described by Stetson (1997). In this technique, each
measurement's weight is altered according to its difference from the
prior estimate of the mean (taken to be the median of the epoch's
measurements for an initial guess), as implemented here according to
the formula:
\begin{equation}
\sigma_i'^2= {{\sigma_i^2} \over{1+({{m_i-\bar{m}} \over{2
\sigma_i}})^2}},
\end{equation}
\noindent where $m_i$ is the $i$th measurement, $\sigma_i$ is the error
estimate in that quantity after the prior iteration, and $\bar{m}$
is the estimate of the mean from the prior iteration; after this
adjustment of the weights, a new determination of the mean is
made.
An estimate of the uncertainty in each epoch's mean magnitude
measurement was obtained from the weighted standard deviation of
the data:
\begin{equation}
\sigma_m^2 = {{\sum{{(m_i-\bar{m})^2} \over{\sigma_i^2}} \over{{\sum{{1} \over{\sigma_i^2}}} (n-1)}}},
\end{equation}
where $n$ is the total number of measurements used in determining
$\bar{m}$. The resulting uncertainty estimates were generally
accurate to 10-20
\% (based upon the median $\chi^2$ of the comparison of each epoch's
magnitude measurements for a star to the mean magnitude obtained from
all F555W measurements for that star).
An additional potential source of photometric errors is the estimation
of the background counts underlying the star (due to unresolved stars,
H II regions, etc.). ALLFRAME estimates that background level by
taking the median number of counts from pixels within some annulus
about the star whose magnitude is being measured. Initially, our
studies were done using an annulus from 3 to 20 pixels in radius from
the stars; we later performed photometry using background annuli from
3 to 10 pixels and from 3 to 6 pixels. For $F555W$ observations of
faint stars with well-determined photometry, the mean change in
epochal magnitudes was 0.000 $\pm 0.001$ mag, the RMS 0.10 mag, and
the root median square difference (also known as the probable error) 0.047 mag when photometry was done
with a 3-10 pixel radius sky annulus instead of 3-6. For $F814W$, the
corresponding numbers were 0.000 $\pm$ 0.003 mag, 0.18 mag, and 0.079
mag. Somewhat larger differences resulted from changing from a 3-10
pixel background annulus to 3-20, though increasing the background
region does reduce the scatter among the magnitude measurements for a
given star. Therefore, for the mean magnitudes presented here, we have
adopted the 3-10 pixel background level and included the probable magnitude error within the uncertainty estimate for each star's magnitude.
We found that using the ALLFRAME error estimates for weighting when
averaging magnitudes for a given star yielded a systematic bias
towards the brighter measurements. This bias is minimal when errors
are small, but is several tenths of a magnitude for the faintest
stars. We have chosen to perform averaging of magnitudes rather than
fluxes as it yielded a lower scatter of epochally averaged magnitudes
in tests of both artificial and actual stars, and our ability to
discriminate variations in brightness from the effects of magnitude
measurement errors was a major limiting factor in this work; a much
smaller but significant bias in the opposite sense was also found for
flux averaging. A comparison of the averaged $F555W$ magnitudes for
stars on Chip 1 to their unbiased median magnitude measurements may be
found in Figure 2, along with a functional fit to the bias (here and
in the remainder of the paper, $F555W$ and $F814W$ will refer to magnitudes
obtained by combining ALLFRAME measurements with an appropriate zero
point; no aperture or bias corrections have been applied to them. $V$
and $I$ will be used to refer to fully corrected magnitudes on the
Johnson and Kron-Cousins systems, respectively). Such functional fits
were used in a Brent's method-based algorithm (cf. Press et al. 1992)
to remove the biases in mean $V$ and $I$ magnitudes before color
measurements or comparison to Cepheid P-L relations. Any biases due
to averaging procedures should be corrected via this method. The
expected error in the amount of the bias correction due to errors
in measuring a star's mean magnitude is much less than the width of
the P-L relation in both $V$ and $I$ ($\sim$ 0.04 mag for typical
candidate Cepheids in our dataset), and thus should have no effect on
our final results.
\subsection{Cepheid Identification}
NGC~4603 is the most distant galaxy for which a Cepheid search has
been attempted; the required photometry presented a considerable
challenge. Because the errors in the magnitude measurements for each
epoch were a substantial fraction of typical Cepheid amplitudes and
because of the limited number of epochs available, common techniques
for identifying variables (see, e.g., Rawson et al. 1997 and
references therein) proved to be of limited utility; for instance, the
phase dispersion minimization method, which requires binning the
observations in phase, is hardly optimal for noisy datasets with such
a limited number of observations (Stellingwerf 1978). We instead have
adopted an alternative approach loosely based on that described in
Stetson 1996.
\vbox{%
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=8.9cm
\epsffile{f2.eps}}
\begin{small}
\figcaption{\small A plot of the difference between the unbiased
median magnitude measurement for a star and the biased epochal average
measurements. It is readily apparent that this bias is greater at
fainter magnitudes. The solid line traces the median bias for stars
in 0.1 mag wide bins, with error bars corresponding to the standard
error of the mean for each bin; the dashed line is a regression fit to
that data, showing that the bias is well represented by a power law in
the actual (as opposed to measured) flux. So long as there exists a
one-to-one correspondence between the unbiased actual and biased
measured magnitudes (which is true for measured $F555W < $ 28.125,
$F814W < $ 27.166), we may use this fit relation to correct for the
bias.}
\end{small}
\end{center}}
The computing power of modern workstations is now sufficient that we
could attempt to fit the photometry for every well-observed star to a
grid of model Cepheid light curves (taken from Stetson 1996) with a
wide range of periods and phases (in general, we sampled the period in
1 day increments and phase in increments of 0.025 for our variable
search). This reduces the problem to a set of linear regressions to
obtain mean magnitude and amplitude, a quite rapid procedure. By
minimizing $\chi^2$ on this grid, we obtain an estimate of the most
appropriate Cepheid light-curve parameters for a given star.
Nonvariable stars emerge from this fitting process with low
amplitudes, typically substantially smaller than the amplitude error
estimates resulting from the procedure; they can be rejected on this
basis. For variables, the width of the minimum in the variation of
$\chi^2$ with period allows us to estimate our uncertainty in that
parameter for a given star. We also confirmed our light-curve fits by
performing a nonlinear $\chi^2$-minimization fit to the data for
suspected variables with our best-fitting parameters as initial
guesses. This generally resulted in minimal changes in parameters,
indicating that our grid was sufficiently fine.
There are complications for longer-period variables ($> 40$d), for
which multiple deep minima in $\chi^2$ appear due to aliasing.
However, our Monte Carlo analysis (see $\S 3.3.2$) indicates that we
still determine the periods of such Cepheids to 10-20 \% accuracy
(with the period errors then dominated by the spacing between the
minima, reflecting the possibility that the deepest $\chi^2$
minimum occurs at an alias --- typically the nearest one --- of the
actual period, as reflected in Figure 1).
Stetson (1996) also defines model Cepheid $I$-band light curves
based upon the same parameters as those for $V$. Thus, once
a $V$-band fit is obtained, two determinations of the mean $I$
magnitude for a star can be made by combining our two epochs'
magnitude measurements and the expected $I$ variation at the phase
of those measurements (a method not unlike that described in
Sandage et al. 1997). We used a weighted mean of these two
determinations to estimate the mean $I$ magnitude for our
variables.
\subsection{Simulations of Cepheid Detection Rates and Expected
Errors}
Because the Cepheids we are looking for are so faint, it is critical
to confirm our ability to unambiguously detect such stars and to limit
contamination of our sample of Cepheids with nonvariable stars. We
therefore performed our variable search on two sets of artificial
photometric data, one consisting of intrinsically nonvariable stars
and one of stars changing in brightness (before measurement errors)
according to template Cepheid light curves. For each star in one of
these datasets, artificial magnitude measurements were made according
to the actual timing of the HST visits. To account for the
possibility of non-Gaussian distributions of errors, these constant or
varying light curves were modified by numbers selected randomly from
the set of actual deviations of $F555W$ or $F814W$ magnitude
measurements of stars in a given magnitude range from their overall
robustly determined mean magnitude (which, having been found from 48
or 12 magnitude measurements, respectively, were much more accurate
than a single-frame measurement). To retain the information on a
measurement's quality present in the ALLFRAME error estimates, each
frame's magnitude measurement in the artificial datasets was assigned
the magnitude uncertainty estimate from the appropriate star and frame
number for the measurement error applied. This analysis was performed
for stars in ten 0.2 magnitude wide ranges, equivalent to $F555W$
magnitudes from 26.3-26.5 to 27.7-27.9.
\subsubsection{False Positives}
Attempting to find variables in our fake photometry of nonvariable
stars generates candidate ``variables'' that mimic real Cepheids,
hereafter referred to as ``false positives.'' The rate of these
misidentifications in the NGC~4603 dataset is such that any reasonable
list of candidate Cepheids we may produce will be contaminated with
nonvariable stars. However, using our simulations, we have been able
to find a number of criteria that can help reject such candidates.
Some results of these simulations are plotted in Figures 3-5.
Foremost, the majority of the false positives possess low amplitudes
($<$0.6 magnitudes peak-to-peak in the principal Fourier component,
the form of amplitude measured by our template fitting technique), as
illustrated by Figure 4, so excluding low-amplitude variables
eliminates many of them. We also exclude stars with low amplitudes
compared to their statistical error estimates from least-squares
fitting. A further test that proved very useful was to require all
candidate Cepheids to have at least four data points more than 1.2
$\sigma$ away from their robustly determined overall mean magnitude;
nonvariable stars rarely possessed that many deviating points. This is
effectively a test for a non-Gaussian distribution of magnitude
measurements (a characteristic of Cepheid light curves) that is
resistant to a small number of outliers. Another helpful restriction
was ruling out very short period ($< 24 $d) candidates, as those were
far more likely to be false positives than real Cepheids (due to the
increasing ability to make a given light curve match given magnitude
variations with some choice of phase at shorter periods). All results
discussed in this paper utilize variable-finding routines that perform
all of these tests. The number of nonvariable stars which survive our
variability criteria is fairly low; as shown in Figure 3, on Chip 1
(the Planetary Camera, which has the deepest effective photometry) we
find that $\sim 0.5$ \% of all faint stars with $F555W \simeq 27$ (but
up to 6 \% by $F555W=27.6$) may be misclassified as Cepheids in our
analysis.
\vbox{%
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=7.5cm
\epsffile{f3.eps}}
\begin{tiny}
\figcaption{\small Results of simulations for the rate at
which false positives occur for stars on Chip 1 (the Planetary
Camera) as a function of $F555W$ magnitude using our variability criteria, which have excluded the great majority of such misidentified stars. In
this and all following figures, the dotted line indicates the fit
used in obtaining maximum likelihood estimates of distance (see $\S 3.1$).}
\end{tiny}
\end{center}}
Given the large numbers of faint stars in our dataset, it is likely
that our list of candidate Cepheids contains many which are actually
nonvariable. There were roughly 3000 stars with well-determined
photometry (i.e., magnitude measurements on $>$ 90\% of all frames) on
Chip 3, which has the most stars found; there are roughly 2100 such
stars on Chip 1 (the Planetary Camera). We searched for Cepheids
among these. Integrating the false positive rate over our observed
magnitude distributions, we expect 34.0 $\pm ~5.8$ on Chip 1, and 56.9
$\pm ~7.5$ on Chip 3. In contrast, 61 stars on Chip 1 (all of $F555W$
magnitude 27 or\\
\vbox{%
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=7.5cm
\epsffile{f4.eps}}
\begin{small}
\figcaption{\small Results of simulations for the distribution
of false positives in amplitude for magnitudes typical of our
Cepheid candidates. Stars with measured amplitude below 0.60 mag
were rejected as Cepheid candidates.}
\end{small}
\end{center}}
fainter) and 69 on Chip 3 passed all our Cepheid
detection tests. Therefore, we concluded that the set of putative
variables on Chip 3 is too contaminated to yield useful information,
and we have concentrated on the candidate Cepheids on Chip 1 for further
analysis.
\subsubsection{Artificial Cepheids}
To determine our ability to detect any variable stars present in our
dataset, we generated data with realistic photometric errors determined as
described above applied to analytically defined Cepheid light curves
(Stetson 1996) with randomly selected periods, amplitudes, and phases. These
``artificial Cepheid'' Monte Carlo simulations yielded encouraging
results. On both the Wide Field and Planetary Camera chips, 45-70 \%
(depending upon input parameters; the recovery rate was substantially
less than this for candidates with input amplitudes below 0.6 mag, as
should be expected given our variability criteria) of those Cepheids
with mean $F555W \ltsim 27.5$ passed our tests, with probable
magnitude measurement errors of $\ltsim 0.1$ mag and period errors of
$\sim 10 \%$, quite comparable to the uncertainty estimates from our variable
search routines. Some of the results of these simulations are
presented in Figures 6-9.
\section{Results}
A number of potential Cepheid variables on Chip 1 with well-
determined parameters were found. Light curves for some of these candidates
are shown in Figure 10. Their properties are summarized in Table 2. Epochal photometry and light curves for all candidate Cepheids are available via WWW.
\footnote{\texttt http:$\slash\slash$www.astro.berkeley.edu/$\sim$marc/n4603/}
\vbox{%
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=7.5cm
\epsffile{f5.eps}}
\begin{small}
\figcaption{\small Results of simulations for the distribution
of false positives in period, for magnitudes typical of our
Cepheid candidates. Stars with measured period less than 24 d or
greater than 60 d were rejected as Cepheid candidates.}
\end{small}
\end{center}}
The $F555W$ and $F814W$ mean magnitudes of the variables determined
from the chi-squared minimization were converted to Johnson $V$ and
Kron-Cousins $I$ using equations from Hill et al. (1998):
\begin{eqnarray}
V = F555W-25&-&0.052(V-I) \\
\nonumber &+&0.027(V-I)^2+22.510 \\
I=F814W-25&-&0.063(V-I) \\
\nonumber &+&0.025(V-I)^2+21.616, \\
\nonumber \end{eqnarray}
where $F555W$ and $F814W$ are the measured ALLFRAME magnitudes for
the corresponding filters. Fixed aperture corrections of $-0.17
\pm 0.01$ magnitudes each, determined based on those obtained in
prior Key Project ALLFRAME analyses for the PC (Graham et al. 1998), were
also applied when obtaining the $V$ and $I$ magnitudes.
\subsection{Maximum Likelihood Analysis}
In order to extract as much of the information available from our set
of candidate Cepheids as we can despite the presence of false
positives, we have performed an extensive maximum likelihood analysis
to determine the distance modulus of NGC~4603. This required
knowledge of our variable detection rates and the errors in measuring
the period and magnitude of actual Cepheids, in addition to the
distribution in period, magnitude, and amplitude of false positives;
these could all be obtained from our Monte Carlo simulations
(q.v. above). To perform the maximum likelihood analysis, we also
required some knowledge of the distribution in period of actual
Cepheids; this was
\vbox{%
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=7.5cm
\epsffile{f6.eps}}
\begin{small}
\figcaption{\small Results of simulations for the rate at which our algorithms detect Cepheid variables of a given $F555W$ magnitude. The data are divided into subgroups according to the periods of the simulated Cepheids
(with the largest symbols used for the average detection rate for
those stars with the longest periods of variation, and the smallest
the shortest). The dependences of the variable detection rate upon
period and amplitude were complex and nonseparable, requiring us to
interpolate upon a grid of simulation results in our maximum
likelihood analysis.}
\end{small}
\end{center}}
found through a power-law fit to the long-period
tail of the distribution of Large Magellanic Cloud (LMC) Cepheids in
Alcock et al. 1999 to be roughly proportional to the -2.0 power of
period (defining the parameter $\alpha$ used below; i.e., we have
adopted a differential distribution of Cepheids in period of the form
$N(P_R) dP_R \propto P_R^{\alpha} dP_R$). Even violently changing
this assumption (changing $\alpha$ by $\pm$ 1) led to changes in the
derived distance modulus of less than 0.10 mag. For the purpose of
this analysis, we adopt the LMC Cepheid Period-Luminosity relations of
Madore \& Freedman (1991) (and, for the likelihood analysis, the
dispersions of LMC Cepheids about that relation) which have been used
by the Key Project on the Extragalactic Distance Scale.
There are two distribution functions required for this
analysis, labelled hereafter as $f_{real}$ and $f_{false}$. These
represent the probability that a particular star is a real Cepheid
and detected with given properties, or a nonvariable star and
identified as a Cepheid with those properties, respectively.
Based upon the results of our simulations, the former is defined
as a function of observed period $P$, magnitude $m$, and amplitude
$A$, and of the given distance modulus $m-M$ as
\begin{eqnarray}
f_{real}(m,P,A | m-M) \sim p_{detect}(m,P,A) {1 \over 2 \pi \sigma_{m}
\sigma_{P}} \\
\nonumber \times \int_{P_{min}}^{P_{max}}{ e^{-{(P-P_r)^2 \over 2
\sigma_P^2}} e^{-{(m-m_r)^2 \over 2 \sigma_m^2}} P_r^{\alpha}
dP_r},
\end{eqnarray}
where $p_{detect}(m,P,A)$ is the probability of our detecting a
Cepheid that has a given observed magnitude, period, and amplitude,
$P_r$ is the actual, as opposed to
\vbox{%
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=7.5cm
\epsffile{f7.eps}}
\begin{small}
\figcaption{\small Results of simulations for the probable
error in measuring the mean $F555W$ magnitude of a Cepheid as a
function of $F555W$ magnitude. No trend in this quantity is seen
either with the period or amplitude of the Cepheid's variation; in
this plot, as in Figure 6, symbols of larger size represent errors for
longer-period variables. In our maximum likelihood analyses, probable
errors are multiplied by an appropriate correction factor, 1.48260, to
yield the corresponding Gaussian $\sigma$.}
\end{small}
\end{center}}
observed, period of a Cepheid,
$m_r$ is the ideal magnitude of a Cepheid for a given distance modulus
and $P_r$ (from the Madore \& Freedman P-L relation), and $\alpha$ is
treated as a constant parameter for the maximum likelihood analysis
describing the distribution in period of actual Cepheids. In our
Monte Carlo analysis, $\sigma_{P}$ proved to be a complex function of
period, amplitude, and magnitude, while $\sigma_m$ was significantly
dependent only on magnitude (as applied in the maximum likelihood
analysis, $\sigma_m$ has added to it in quadrature contributions from
the dispersion of the P-L relation and estimates of magnitude
measurement errors due to background subtraction and bias correction
uncertainties). Because of their complicated dependence on all
possible variables, values of $v$ and $\sigma_P$ were obtained by
interpolating within a $9 \times 9 \times 9$ grid in period,
amplitude, and magnitude containing the results of simulations for
these quantities. The values of $\sigma_m$ were taken from
least-squares fits of empirically chosen functions to the Monte Carlo
results. Once a distance modulus is chosen, $f_{real}$ is normalized
to make the expectation value of the number of Cepheids existing in our dataset unity:
\begin{equation}
\sum_i{n_i \int{f_{real}(m_i,P | m-M) dP=1}},
\end{equation}
where $m_i$ is the mean magnitude in a bin (0.04 mag wide in our analysis) and $n_i$ is the
number of observed stars with good photometry in that bin. The actual
number
\vbox{%
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=7.5cm
\epsffile{f8.eps}}
\begin{small}
\figcaption{\small Results of simulations for the probable
error in measuring the mean $F814W$ magnitude of a Cepheid as a function of
$F555W$ magnitude. No trend in this quantity is seen with the
period or amplitude of the Cepheid's variation.}
\end{small}
\end{center}}
of Cepheids in the data will then be a parameter whose value is
determined by the likelihood analysis. This integral is performed
numerically in 1 day increments over the range of periods accepted for
candidates, 24-60 days. This normalization may not be perfect,
particularly for the $I$ analysis, so much more significance should be
ascribed to, e.g., results for the difference of the number of
Cepheids from zero than to the exact number of Cepheids found.
The chance that a given unvarying star is selected as a candidate
variable of given properties, $f_{false}$, was found via our
simulations to be proportional to a power law in period and a Gaussian
(of zero mean) in amplitude:
\begin{equation}
f_{false}(m,P,A) \sim {{1} \over{\sqrt{2 \pi \sigma_A}}} e^{-{{A^2}
\over{2\sigma_A^2}}} P^{\beta};
\end{equation}
in our Monte Carlo simulations, $\sigma_A$ proved to be a function of
the magnitude alone and $\beta$ a constant. Because these parameters
are independent of period and amplitude, $f_{false}$ may be
integrated over these variables analytically. This distribution was
then normalized such that its integral over the possible periods and
amplitudes for candidates was 1; for an individual candidate, it must
be multiplied by the overall rate of false positives at a given
magnitude, $r(m)$, to yield the probability that that star is a false
positive.
The functional fits to the parameters required by the maximum
likelihood analysis used were:
\begin{equation}
\sigma_{m_V}(m_V)=0.1 \times 10^{-0.2824(28.18-m_V)} $$
$$ \sigma_{m_I}(m_V)= 0.2 \times 10^{-0.2696(28.09-m_V)} $$
$$ \sigma_A(m_V) = \max (0.08,-2.013+0.0791 m_V) $$
$$ \beta = -1.82 $$
$$ r(m_V) = \max (0.05350 \times (m_V-26.6)^{2.714},0),
\end{equation}
where $m_V$ is the $F555W$ magnitude of a given star before
aperture corrections and $m_I$ its corresponding $F814W$
magnitude.
The logarithm of the likelihood is then defined as
\begin{equation}
\ln{\cal{L}}=\sum_{i=1}^{n_{cand}} \ln[ N_{Ceph} f_{real}(m,P,A |
m-M) + r(m) f_{false}(m,P,A)] $$
$$ + \sum_{j=1}^{n_{bin}} n_j \ln\left[ \left(1-N_{Ceph}
f_{real}(m_j | m-M)\right)
\left(1-r(m_j) \right) \right],
\end{equation}
where $n_{cand}$ is the total number of Cepheid candidates, $n_{bin}$
is the number of magnitude bins used, and $N_{Ceph}$ is roughly
equivalent to (and directly proportional to) the number of observed
Cepheids in the dataset, an unknown in the analysis. The first
summation corresponds to the product (before the logarithm) of the
probabilities that our candidate Cepheids will be detected as such
stars with their given properties; the second, the product of the
probabilities that each of our non-candidate stars are not either
detected Cepheids or false positives. The logarithm of the
likelihood, and thus the likelihood itself, is maximized over a grid
in the distance modulus $m-M$ and the number of Cepheids in the
dataset $N_{Ceph}$. Note that for the non-candidates, the
distribution functions have been integrated over period and amplitude.
We have tested our maximum likelihood techniques by applying them to
datasets containing both nonvariable stars (potentially false
positives) and a set of simulated Cepheids with a fixed distance
modulus and a realistic distribution of properties. If the number of
real Cepheids was large enough and NGC~4603 placed near enough that
\vspace{.33 in}
\vbox{%
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=7.5cm
\epsffile{f9.eps}}
\begin{small}
\figcaption{\small Results of simulations for the probable
error in measuring the period of a Cepheid as a function of magnitude.
Like the Cepheid detection rate, this quantity proved to be dependent
on period and amplitude in a complex, nonseparable fashion; therefore,
interpolations of the results of simulations were used for it in our
maximum likelihood analysis. Due to aliasing, the longest-period
Cepheids have the greatest errors in period determination (as can be
seen in Figure 1).}
\end{small}
\end{center}}
a significant number of the input Cepheids are found by the
variable search procedure, then our techniques effectively recovered
the input distance. If those conditions are not met, the
highest-likelihood solutions generally prove to be those in which the
number of Cepheids in the dataset is minimized or the distance modulus
is maximized, i.e., cases in which the chance of observing a Cepheid
would be as small as possible. Such solutions are easy to recognize
and did not occur in our analysis of NGC~4603.
This analysis was performed independently using the $V$ and $I$ mean
magnitudes of our candidates to determine the distance modulus. From
the $V$ analysis, we determine that the hypothesis of no Cepheids
present is excluded at $> 9\sigma$, that $43 \pm~7$ actual Cepheids
are present in our dataset, and that NGC~4603 has a distance modulus
of $33.15_{-0.10}^{+0.11}$ (1$\sigma$ random errors) before correction
for metallicity and dust extinction. The $I$ analysis yields a poorer
constraint, with a Cepheid signal present at $> 7 \sigma$ and an
uncorrected distance modulus measurement of $32.97_{-0.09}^{+0.15}$.
See Figures 11 and 12 for plots of the resulting likelihood contours.
The location of our candidates in the NGC~4603 color-magnitude diagram
is shown in Figure 13. Figures 14 and 15 illustrate the differences
between the distributions of candidate Cepheids in magnitude and color
and those expected for false positives. The excess candidates beyond
the false positives do seem limited in their brightness and colors in
the fashion expected for Cepheids of a variety of periods and
reddenings.
Since we have obtained substantial knowledge about the distribution in
properties of real Cepheids and false positives through our maximum
likelihood analysis, we can estimate the probability that a given
candidate is in fact a Cepheid; the resulting probabilities are listed
in Table 2. $V$ and $I$ period-magnitude plots for the potential
Cepheids we have found in NGC~4603 (containing essentially the
\vspace{.13in}
\setcounter{figure}{10}
\vbox{%
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=7.5cm
\epsffile{f11.eps}}
\begin{small}
\figcaption{\small Results of our maximum-likelihood analysis
using $V$ mean magnitudes of candidate Cepheids to determine the
distance to NGC~4603. The contours represent 1,2,3,4, etc. $\sigma$
limits on the measured parameters. We confirm that Cepheids are
present in our data set at $>9 \sigma$.}
\end{small}
\end{center}}
same
information) may be found in Figures 16 and 17. Those candidates
found to have greater than 50 \% probability of being Cepheids in both
the $V$ and $I$ maximum likelihood analyses have their
simulation-based error bars (as used in the analyses) depicted on the
plots.
Such higher-probability candidates may be used to provide an
illustration of the workings of our maximum likelihood procedure.
These stars should have a relatively high value of $f_{real}$, so they
must agree with the expected magnitude of a Cepheid of the same
measured period given our choice of distance modulus within the
estimated errors. However, they should also have a relatively small
value of $r(m) f_{false}$. Given the strong magnitude dependence, we
expect such stars to be brighter than the typical candidate. Thus, if
we were to calculate the mean distance modulus predicted from the
properties of such stars, we would expect it to be fairly consistent
with but biased low compared to that obtained from the full maximum
likelihood analysis. This is borne out by such a calculation for,
e.g., those stars that have $>80\%$ probability in both the $V$ and
$I$ analyses; they give a value of $32.95 \pm~0.10$ for the $V$
modulus and $32.80 \pm~0.08$ for $I$, 0.20 and 0.17 mag less than
those obtained from the full procedure. The maximum likelihood
technique does not simply determine a distance modulus weighting stars
according to their probability of being Cepheids, but instead
incorporates as much information as possible about how effectively we
can find such stars, minimizing incompleteness/Malmquist-type biases.
\subsection{Uncertainties and Corrections}
In addition to the statistical uncertainties in our measurements of
the distance modulus of NGC~4603, which
\vbox{%
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=7.5cm
\epsffile{f12.eps}}
\begin{small}
\figcaption{\small Results of our maximum-likelihood analysis
using $I$ mean magnitudes of candidate Cepheids to determine the
distance to NGC~4603. The contours represent 1,2,3,4, etc. $\sigma$
limits on the measured parameters. We confirm the detection of
Cepheids in the $I$ band at $>7 \sigma$.}
\end{small}
\end{center}}
\vbox{%
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=7.5cm
\epsffile{f13.eps}}
\begin{small}
\figcaption{\small A color-magnitude diagram for stars on Chip 1. Candidate
Cepheids are indicated by open symbols.}
\end{small}
\end{center}}
were determined by our maximum
likelihood analysis, our results are also subject to a number of
potential sources of systematic error. In this subsection, we will
attempt to estimate the amounts of possible error due to the
calibration of photometry and our analysis techniques, to
uncertainties in the Cepheid P-L relation calibration, and to our
limited knowledge of physical conditions in and towards NGC~4603, and
make whatever well-established corrections possible to our distance
moduli.
Uncertainties in the HST zero point of $\pm 0.05$ magnitudes in $V$ or
$I$ affect our measurements of the NGC 4603 distance in much the same
fashion as Key Project distances (Hill et al. 1998), with one
important difference: because the mean magnitudes of a given star are
not as well determined, we are unable to use a measurement of $E(V-I)$
to measure reddening, so relative zero point errors do not propagate
into our results as they do in the Key Project methodology. A mean
difference of 0.08 mag in $V$ between DoPHOT and non-bias-corrected
ALLFRAME photometry for our candidates was found. This may be due to
differences in the characteristics of any biases that occur when
averaging ALLFRAME and DoPHOT results at these faint magnitudes. We
adopt the bias-corrected ALLFRAME results here and include half this
difference as a potential systematic error in $V$ magnitudes. In the
absence of sufficient DoPHOT comparison photometry in $I$, we consider
a 0.10 mag systematic error to be possible, though substantially
larger than any found in prior studies.
We also must consider
uncertainties in the distance modulus resulting from the maximum
likelihood methodology and the Monte Carlo fits that were used to
define $f_{real}$ and $f_{false}$. Changing the assumed false
positive rate radically (e.g., by 50\%) altered the resulting distance
modulus constraints by at most 0.09 mag in both $V$ and $I$.
Considering also the differences in measured distance modulus
\vbox{%
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=8.9cm
\epsffile{f14.eps}}
\begin{small}
\figcaption{\small Histogram of the $F555W$ magnitude
distribution of candidate Cepheids (solid line) and that expected for
false positives given the distribution in magnitude of observed stars
(dashed line).}
\end{small}
\end{center}}
exhibited when the power-law parameter for the distribution in period
of real Cepheids, $\alpha$, is changed by $\pm 1$, we find potential
systematic errors in the maximum likelihood procedure of 0.14 mag for
$V$ and 0.12 mag for $I$. If we add the corresponding errors in
quadrature, we find that systematic errors in photometry and in our
analysis techniques should be less than 0.15 mag in $V$ and 0.16 mag
in $I$. Because we account for the lower probability of detecting
faint Cepheids via our maximum likelihood technique and fix the slope
of the P-L relation used, the effects of incompleteness bias should be
minimal.
Our results are also subject to possible systematic errors in the
calibration of the Cepheid P-L relation. Indeed, one of the largest
remaining systematic uncertainties in the extragalactic distance scale
is our limited knowledge of the distance to the Large Magellanic
Cloud, which currently provides the fiducial standard Cepheid
calibration. For the Key Project, this uncertainty has been taken to
be $\pm$ 0.13 mag; we adopt this value so that if the distance to the
LMC is better determined in the future our distance determination may
be easily adjusted in concert with theirs. We also adopt the Key
Project's estimate of potential errors within the LMC $V$ and $I$ P-L
calibrations of $\pm 0.05$ magnitudes (see, e.g. Rawson et al. 1997).
A number of potential systematic errors in our distance modulus could
be the result of physical effects. First, as an Sc galaxy or,
alternatively, one with maximum circular velocity of 220 $\,{\rm km\,s^{\scriptscriptstyle -1}}$\
(Giovanelli et al. 1997), we may expect Cepheids in NGC~4603 to
possess substantially higher
\vbox{%
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=8.9cm
\epsffile{f15.eps}}
\begin{small}
\figcaption{\small (upper panel) Histogram of the $V-I$ colors of
candidate Cepheids (solid line) and that expected for false positives
given the color distribution of observed stars (dashed line). (lower
panel) The difference between the two histograms, smoothed with a
Gaussian kernel. The dotted line indicates the typical color expected
for a Cepheid of period 35 d reddened by foreground Galactic
dust as measured by Schlegel, Finkbeiner, \& Davis (1998). The
distribution appears very consistent with Cepheids of a range of
reddenings, with no similar excess of blue stars.}
\end{small}
\end{center}}
metallicity than those in the LMC, by
roughly 0.40 $\pm 0.20$ dex at the radius of the PC field (applying
the results of Zaritsky, Kennicutt, and Huchra 1994 to obtain values
for the typical metallicity and metallicity gradient in NGC~4603).
Using the relation of Kennicutt et al. (1998), we should therefore
expect that our distance modulus is an underestimate by 0.096 $\pm
0.081$ mag. It should be noted that other studies have found larger,
but still statistically equivalent given the error bars, metallicity
effects (Sasselov et al. 1997, Kochanek 1997, Nevalainen \& Roos
1998), while theoretical calculations predict effects that are minimal
or opposite in sign (Alibert et al. 1999, Musella 1999). We therefore
make no correction, and consider the entire 0.096 mag to be a
potential systematic error.
Another potential physical effect is extinction by dust along the
line-of-sight to the Cepheids, either within NGC~4603 or our own
Galaxy. Unfortunately, the Centaurus cluster lies behind a region
where substantial emission from Galactic dust has been observed; the
effect of this dust should therefore be quite appreciable. We thus
must correct the distance moduli we have found for Galactic foreground
dust absorption of $A_V$=0.54 $\pm 0.08$ magnitudes ($A_I$=0.33
magnitudes, using a typical Galactic extinction law), taken from the
extinction map of Schlegel, Finkbeiner \& Davis (1998). After
correction for foreground extinction, our data yield $E(V-I)_{internal}=-0.04^{+0.14}_{-0.18}$ (random); we constrain the reddening
due to dust within NGC~4603 only poorly. To place limits on its
effects, we may safely assume that internal dust will yield
E($V-I$)$\geq$ 0, of course; an examination of Key Project papers
studying galaxies of similar inclinations indicates that E(V-I) $<
0.07$ due to internal reddening is also a reasonable
assumption. Conversion to $A_V$ with a typical Galactic extinction law
indicates that we might therefore expect that our $V$ distance modulus
should be reduced by as much as 0.17 mag in correcting for extinction
by dust within NGC~4603, and our $I$ modulus by as much as 0.10 mag.
We thus adopt -0.09 magnitude as an estimate of the possible 1$\sigma$
systematic error from internal extinction in $V$, and -0.05 mag in
$I$.
To estimate the total potential systematic errors, we add possible
errors from physical effects in quadrature to the systematic
uncertainties of the modulus from our photometric and maximum
likelihood techniques and that from the P-L relation calibration,
yielding total systematic
\vbox{%
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=8.9cm
\epsffile{f16.eps}}
\begin{small}
\figcaption{\small The $V$ P-L relation for our candidate
Cepheids. The solid line depicts the LMC P-L relation shifted to
the distance modulus we have obtained; dotted lines indicate the
$2-\sigma$ scatter of LMC Cepheids about that relation. Those candidates
with more than 50\% probability of being Cepheids in both the $V$ and $I$
analyses are plotted with the error bars used for them in
the maximum likelihood analysis (drawn from our simulations).}
\end{small}
\end{center}}
\vbox{%
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=8.9cm
\epsffile{f17.eps}}
\begin{small}
\figcaption{\small The $I$ P-L relation for our candidate
Cepheids. The solid line depicts the LMC P-L relation shifted to the
distance modulus we have obtained; dotted lines indicate the
$2-\sigma$ scatter of LMC Cepheids about that relation. Those
candidates with more than 50\% probability of being Cepheids in both
the $V$ and $I$ analyses are plotted with the error bars used for them
in the maximum likelihood analysis (drawn from our simulations).}
\end{small}
\end{center}}
uncertainties of +0.23/-0.24 magnitudes in
$V$ or $\pm 0.23$ magnitudes in $I$. Combining all effects, we thus
find from the $V$ analysis that NGC~4603 has a distance modulus of
$32.61_{- 0.10}^{+0.11}$ (random, 1 $\sigma$) $^{+0.24}_{-0.25}$
(systematic), corresponding to a distance of $33.3^{+1.7}_{-1.5}$
(random, 1 $\sigma$) $^{+3.8}_{-3.7}$ (systematic) Mpc. The $I$
analysis provides a quite consistent result, yielding a distance
modulus of $32.65_{-0.09}^{+0.15}$ (random, 1 $\sigma$) $\pm 0.24$
(systematic).
\subsection{Implications}
Previous studies have obtained widely differing measurements of the
distances to, and hence peculiar velocities of, Cen30 and Cen45, even
when using the same techniques. Larger $D_n - \sigma$ samples
(e.g. that used in the Mark III catalog [Willick et al. 1997], which
includes 22 galaxies in Cen30 and 9 in Cen45, as opposed to 9 and 4,
respectively, in Faber et al.), for instance, have placed Cen30 as far
as or even $behind$ Cen45, with peculiar velocities of -110 $\,{\rm km\,s^{\scriptscriptstyle -1}}$\ and
+1515 $\,{\rm km\,s^{\scriptscriptstyle -1}}$\, respectively, in great contrast to the earlier results.
The hypothesis that Cen30 and Cen45 may lie at the same distance was
first advanced by Lucey, Currie and Dickens (1986a) on the basis of a
number of relative distance measures. It appears from some studies as
though the Centaurus cluster may be in the midst of a substantial
merger, with Cen45 falling into Cen30 and acquiring a rapid velocity
thereby. On the other hand, some distance measurements, particularly
those considered by Lynden-Bell et al. (1988), imply that Cen30 and Cen45
moving rapidly relative to the Local Group.
A flow of the center of mass of the Cen30/Cen45 system with
such high speed -- faster than the motion of the Local Group itself --
would suggest the existence of a substantial attracting mass. From the
results of the first $D_n-\sigma$ studies, Lynden-Bell et al. (1988)
hypothesized the existence of a ``Great Attractor,'' a large
concentration of matter lying beyond the Centaurus cluster. However,
neither optically nor $IRAS$-selected samples of galaxies have revealed
regions of overdensity sufficient to explain these motions. In fact,
according to redshift surveys, the Centaurus cluster itself, when
combined with Hydra and Pavo-Indus-Telescopium on the other side of
the galactic plane, should constitute the major local attractive
point. Centaurus should therefore be approximately at rest in the
Cosmic Microwave Background reference frame, and the bulk of the
motion of the Local Group driven by its mass overdensity. In the
Local Group frame one then expects to observe negative peculiar
velocities in the direction of Centaurus, as there is expected to be a
strong reflex dipole pattern from the motion of the Local Group
itself. One possible explanation for the discrepancy between
predicted and observed flows in this region has been provided by
Guzman and Lucey (1993), who have suggested that $D_n - \sigma$
distances can be compromised by age effects and that the large outflow
of Centaurus is potentially suspect as a result; however, there is no
particular reason to expect that galaxies in the Centaurus region
should be younger than others in our neighborhood. If there is in fact
only a very weak reflex signature in the velocity of the Centaurus
cluster, the density parameter of the Universe must be very low, too
low to explain the infall pattern around the Virgo supercluster.
The Cepheid distance measurement we have obtained may be used to set
limits on such a flow. Our result is most easily compared to studies
using other distance indicators and a peculiar velocity determination is
most straightforwardly made by converting to velocity distance. This
may be accomplished by multiplying the distance obtained by an
appropriate value for Hubble's Constant based upon the same
calibration; we use the Key Project's most recent estimate for
Hubble's Constant based upon Cepheid data analyzed similarly to that
presented here, 72 $\pm 5$ (random) $\pm 12$ (systematic) $\,{\rm km\,s^{\scriptscriptstyle -1}}$\
$\mathrm{Mpc}^{-1}$ (Madore et al. 1999). We then determine a velocity
distance for NGC~4603 of 2395 $\pm 306$ (random) $\pm 281$
(systematic) $\,{\rm km\,s^{\scriptscriptstyle -1}}$. Note that this value should not be altered by any
recalibrations of the zero point of the Cepheid distance scale because
our distance measurements and those of the Key Project would be
affected in the same way.
A variety of other estimates of the velocity distance of of NGC~4603
and of Cen30 as a whole are presented in Table 3. To allow more
effective comparison, the presumed velocity of Cen30 in the Local
Group frame(``$cz_{Cen30,LG}$'') and number of galaxies included in
each study are also listed; each of the $D_n-\sigma$ samples in the table
includes the preceding work as a subset.
Our result agrees well with estimates of the distance to Cen30 based
on global analyses of the properties of cluster galaxies. Jerjen \&
Tammann (1997), for instance, find from an analysis of galaxy
luminosity functions that Cen30 is 1.63 $\pm$ 0.15 mag beyond Virgo.
Taking the Virgo distance modulus to be 31.07 $\pm$ 0.07 (random; from
Freedman et al. 1998, excluding the 3.9$\sigma$ outlier NGC 4639 from
the average), this yields distance modulus 32.70 $\pm$ 0.16 (random).
Studies of surface brightness fluctuations, too, find distances
consistent with that we have obtained for NGC~4603; recent results of
Tonry et al. (1999) yield a mean distance of $2524 \pm 435$ (random)
for 8 galaxies in Cen30.
Our result may also be compared to peculiar velocity predictions based
on the gravity field inferred from the full sky $IRAS$ survey or from
surveys of optically selected galaxies (Nusser and Davis 1995). For
example, using the gravity field derived for the $IRAS$ 1.2 Jy survey
and assuming $\beta = 0.5$ leads to a predicted peculiar velocity
$v_p$ (in the LG frame) of 14 $\,{\rm km\,s^{\scriptscriptstyle -1}}$\ versus a Cepheid inferred $v_p$ of
-74 $\pm 306$ (random) $\,{\rm km\,s^{\scriptscriptstyle -1}}$\ if the redshift of NGC~4603 is left at
its observed value, $cz_{lg} = 2321$ $\,{\rm km\,s^{\scriptscriptstyle -1}}$, which should be appropriate
if it is in fact a field galaxy. If we instead compare to the central
redshift of Cen30, $\approx 2807$ $\,{\rm km\,s^{\scriptscriptstyle -1}}$\ in the Local Group frame
(Lucey et al. 1986a), then its predicted $v_p$ is -90 $\,{\rm km\,s^{\scriptscriptstyle -1}}$, while the
Cepheid distance would imply $v_p = 412$ $\,{\rm km\,s^{\scriptscriptstyle -1}}$. The predicted and
observed peculiar velocities disagree in this case by $\sim 1.2
\sigma$.
Since we only have been able to perform a Cepheid distance analysis
for one galaxy, we cannot claim to have established unambiguously the
distance to the Centaurus cluster; while some studies have included
NGC~4603 in Cen30, for instance, others have not. Indeed, as
illustrated in Table 3, the location of Cen30 itself in redshift
space, not only real space, has varied substantially from analysis to
analysis, reflecting in no small part the large velocity dispersion
and limited numbers of cluster spirals (Stein et al. 1997). It is
worthy of consideration, though, that those studies that do exclude
this galaxy from Cen30 place it nearer to us than the cluster itself
(as in Willick et al. 1997, though the groupings used for Mark III
spirals tend to place Cen30 at a substantially higher velocity than
that found in other studies), lending support to the higher distance
estimates for the cluster. At worst, our distance measurement should
provide a lower limit on the distance to Cen30, and hence an upper
limit on its peculiar velocity.
Our results are most easily reconciled with those of recent
velocity-distance calibrated studies if NGC~4603 is treated as an
object in the foreground of the Centaurus cluster. That is a rather
reasonable scenario; previous studies (Bernstein et al. 1994, Willick
et al. 1995, Willick 1999) have found that the Tully-Fisher distances
of what are nominally cluster spirals correlate well with their
(rather than their clusters') redshifts. As a galaxy with a
Tully-Fisher distance, NGC~4603 should be subject to the same
selection effects. We note that the velocity distances determined
from the two largest samples of Cen30 galaxies listed in Table 3 are
in excellent agreement with each other though those distances were
obtained via different methods and calibrated separately, and those
measurements agree well with the $IRAS$ 1.2 Jy survey-predicted peculiar
velocity for Cen30. Our distance measurement for NGC~4603 is in good
accord with a variety of Tully-Fisher measurements of the distance for
that galaxy but agrees more poorly with the best measurements of the
distance to Cen30 as a whole. Under very reasonable assumptions, we
may conclude that Tully-Fisher distances, and therefore (based on
their consistency for Cen30) those obtained via the $D_n-\sigma$
technique as well, agree with the Cepheid distance scale and
$IRAS$-predicted peculiar velocities to at least as far away as the
Centaurus cluster.
The rough agreement of the results of this analysis with other studies
of the Centaurus cluster is encouraging. For a firmly established
Cepheid distance to Centaurus, a similar study would have to be
performed on more galaxies, preferably including ones that show more
definitive evidence of location in the cluster core (e.g. stripping of
galactic gas) ensuring that members of Cen30 are observed. However,
the substantial resources in HST time required with current
instrumentation and the extremely extensive data analysis effort
needed to produce a convincing result means that such efforts should
most likely await the installation of the Advanced Camera for
Surveys. Finding Cepheids at the distance of the Centaurus cluster is
currently possible, but difficult indeed.
-----------------
\acknowledgments
\centerline {\bf Acknowledgements}
We would like to thank Jay Anderson for his efforts to provide
secondary photometry for this study and Jeff Willick for his
assistance in interpreting Mark III data. We also would like to thank
our program coordinators at STsCI, Doug van Orsow and Christian Ready,
for their assistance, and the anonymous referee for helpful comments.
This work was supported by NASA grants GO-06439 and GO-07507 from the
Space Telescope Science Institute (operated by AURA, Inc. under NASA
contract NAS 5-26555). J. A. N. acknowledges the support of a
National Science Foundation Fellowship and the Berkeley Fellowship.
|
\section{Introduction}
Let $f:U \to {\rm Spec \,}(K)$ be a smooth open curve
over a field $K\supset k$, where $k$ is an algebraically closed field of
characteristic zero. Let
$\nabla : L \to L\otimes \Omega^1_{U/k}$ be a possibly irregular
absolutely integrable (or vertical, see definition \ref{def2.15})
connection on a line bundle $L$.
The Riemann-Roch problem in this context is
to describe characteristic classes for the relative de Rham cohomology
$Rf_{*}(L\otimes\Omega^*_{U/K})$ as a (virtual) vector space over
$K$ with an integrable connection, in terms of data on $U$. The $0$-th
characteristic class, the Euler characteristic $\dim R^0 - \dim R^1$,
is well-known to be given by
\eq{2-2g-n-\sum_i \max(0,m_i-1)
}{1.1}
where $g$
is the genus of the complete curve $C$, $n$ is the number of missing
points, and $m_i$ is the order of the polar part of the connection at
the $i$-th missing point. The purpose of this article is to give a
formula for the first characteristic class, which is the determinant
of the Gau\ss-Manin connection on the relative de Rham cohomology of
the line bundle,
\eq{\Big(\det(Rf_*(L\otimes\Omega^*_{U/K})), \det\nabla_{GM}\Big).
}{1.2}
When $U=C$, so the connection has no poles, the formula given in
\cite{BE} is
\eq{\Big(\det(Rf_*(L\otimes\Omega^*_{C/K})), \nabla_{GM}\Big) =
-f_*((L,\nabla)\cdot
c_1(\Omega^1_{C/K})).
}{1.3}
Concretely, if one has $c_i \in C(K)$ with $\sum c_i$ a $0$-cycle in
the linear series representing
$\Omega^1_{C/K}$, then the determinant is given by restricting $L$ with its
connection to each $c_i$ and then tensoring the resulting lines
with connection together.
When the connection $\nabla$ has at worst regular singular points at
the points in $D:=C-U$ there is an analogous formula using linear
series given by
divisors of rational sections $s$ of $\Omega^1_{C/K}(D)$ satisfying the rigidity
condition ${\rm res}_D(s) = 1$. Indeed, these formulas are valid also
for higher rank connections. One takes the determinant at zeroes and poles of
$s$.
In the case of irregular singular points, a
similar formula is possible, but the
rigidification taken must depend on the polar part of the connection.
Let $({\mathcal L}, \nabla)$ be an extension of $(L,
\nabla)$ to $C$, ${\mathcal D}= \sum_i m_iD_i$ be a divisor with
multiplicities $m_i \ge 1$ supported in $C-U$ such that the relative
connection
\eq{\nabla_{/K}:{\mathcal L} \to {\mathcal L}\otimes\Omega^1_{C/K}({\mathcal D})
}{1.4}
yields a complex quasiisomorphic to
$j_*L \to j_*L\otimes \Omega^1_{(C-D)/K}$ and has poles
at all points $D_i$. Then
$\nabla_{/K}$
does not factor through
\eq{\nabla_{/K}:{\mathcal L} \to {\mathcal L}\otimes\Omega^1_{C/K}({\mathcal D}-D_i)}{1.5}
for any $i$.
Writing ${\mathcal D}$ also for
the artinian subscheme of $C$ determined by ${\mathcal D}$, this implies
that $\nabla_{/K}$ induces a {\it function linear isomorphism}
\eq{\nabla|_{\mathcal D} : {\mathcal L}|_{\mathcal D} \stackrel{\cong}{\to} {\mathcal L}\otimes
\Omega^1_{C/K}({\mathcal D})|_{\mathcal D}}{1.6}
Because these maps are function linear, we may cancel the lines
${\mathcal L}|_{\mathcal D}$ and deduce canonical elements ${\rm triv}_\nabla \in
\Omega^1_{C/K}({\mathcal D})|_{\mathcal D}$. We view ${\rm triv}_\nabla$ as a trivialization of
$\Omega^1_{C/K}({\mathcal D})$ along ${\mathcal D}$. It is known (\cite{EV}, Appendix B)
that the coboundary of
${\rm triv}_\nabla$ in $H^1(C,\omega_{C/K}) \cong K$ is given by the
degree of ${\mathcal L}$. Our main result is:
\begin{thm} Let notation be as above. Assume all $D_i$ are defined
over $K$ and some $m_i \geq 2$. Because we are only concerned with the
cohomology over $X-{\mathcal D}$, we can take
${\mathcal L}$ of degree 0 so that the element ${\rm triv}_\nabla$ can be lifted
to $ H^0(C,\Omega^1_{C/K}({\mathcal D}))$. Let $s$ be any such lifting, and write $(s)$
for the divisor of $s$ as a section of $\Omega^1_{C/K}({\mathcal D})$. Note the support
of (s) is disjoint from ${\mathcal D}$. Then
\ml{\det(Rf_*(L\otimes\Omega^*_{U/K})), \det(\nabla_{GM}) \cong \\
-f_*(L\cdot (s))+\tau(L) \in \Omega^1_K/d \log K^*.
}{1.7}
Here $\tau(L)$ is a
$2$-torsion term which can be written
$$ \tau(L) = \sum_i \frac{m_i}{2}d\log(g_{i,0})\in
\frac{1}{2}d\log(K^\times)/d\log(K^\times)
$$
where the connection $\nabla_{/K} =
(g_{i,0}+g_{i,1}z_i+\ldots)dz_i/z_i^{m_i}$ for a local coordinate $z_i$
at $D_i \in {\mathcal D}$.
\end{thm}
Note that $\Omega^1_K/d\log K^\times$ is the group of isomorphism classes
of rank $1$ connections on ${\rm Spec \,}(K)$. Our assumption that points of
${\mathcal D}$ are defined over $K$ is made to avoid complications involving
generalized jacobians in \S 2.
We remark, of course, that part of
our task will be to give a precise definition of the right hand side of
the formula of the theorem. It will appear as a product followed by a
trace, and this definition does not depend on the particular choice of
${\mathcal L}$ above. In particular, this gives a formulation if we don't assume
that
${\mathcal L}$ is of degree 0, and also if we don't assume that
$m_i \geq 2$ for at
least one $i$, that is if $\nabla$ has regular singular points (see theorem
\ref{thm4.5}). The precise general formulation of our theorem is in
\ref{thm4.7}. In the case
that $(s)$ is a sum of $K$-points $c_i$, one may simply take the tensor
product of the lines with connection $L|_{c_i}$.
The right hand side of the formula depends only on the equivalence
class of $(s)$ in a generalized Picard (or divisor class) group of
line bundles with trivializations along ${\mathcal D}$. Thus, by analogy with
\eqref{1.3}, it is natural to write formula \eqref{1.7} in the form
\ml{\big((\det(Rf_*(L\otimes\Omega^*_{U/K})), \det(\nabla_{GM})\Big)
\cong \\
f_*\Big(L\cdot c_1(\Omega^1_{C/K}({\mathcal D}),{\rm triv}_\nabla))\Big)^{-1} + \tau(L).
}{1.8}
The classical Riemann-Roch pattern begins to break down in that the
characteristic class $c_1(\Omega^1_{C/K}({\mathcal D}),{\rm triv}_\nabla)$ depends on more
than just the geometry of $f:U \to {\rm Spec \,}(K)$.
This reflects the fact that the de Rham cohomology of an irregular
connection depends on more than topology.
There is an analogy here with the case of $\ell$-adic sheaves. If ${\mathcal E}$ is an
unramified $\ell$-adic sheaf on a complete curve $C$ over a
finite field ${\mathbb F}_q$,
then the global epsilon factor is given by
$$(-F_q|\det(H^*_{\text{\'et}}(C_{\overline{{\mathbb F}}},{\mathcal E}))) = \det({\mathcal E})^{-1}\cdot
c_1(\Omega^1_{C/{\mathbb F}_q}).
$$
The basic result in the ramified $\ell$-adic case (\cite{L}) is that the global
epsilon factor can be written as a product of local terms corresponding to
points on the curve where the sheaf ramifies or where a chosen meromorphic
$1$-form has zeroes or poles. We suspect formula \eqref{1.8} is
analogous to a classical formula for Gau\ss\ sums
$$g(c,\psi) = \sum_{a\in ({\mathcal O}/\frak f)^\times} c(a)\psi(a)
$$
where $\frak f \subset {\mathcal O}$ is an ideal in the ring of integers in a
local field, $c$ (resp. $\psi$) is a character of $({\mathcal O}/\frak
f)^\times$ (resp. $({\mathcal O}/\frak f)^+$), and both $c$ and $\psi$ have
conductor $\frak f$. If the residue field of ${\mathcal O}$ has $q$ elements with
$q$ odd, one finds
$$g(\epsilon,\psi) = \begin{cases}q^nc(x) & \frak f=\frak
m^{2n},\ n\ge 1 \\
q^nc(x)\sigma & \frak f=\frak
m^{2n+1},\ n\ge 1. \end{cases}
$$
In this formula $x\in {\mathcal O}/\frak f$ is a suitable
point, $\sigma=\zeta\sigma_0$ with
$\zeta^q=1$ and
$\sigma_0^2 =
\Big(\frac{-1}{{\mathbb F}_q}\Big)$. (Here $\sigma_0$ is a quadratic Gau\ss\ sum.)
Our proof follows the main idea of Deligne
\cite{De}. For computing the $\epsilon$-factor associated to a rank one
Galois representation on a curve, he expresses the determinant
of the cohomology as the cohomology on a symmetric
product of $(C-D)$ and reduces the computation to
the geometry of the generalized jacobian.
In the geometric situation one is further able
to express the determinant
Gau\ss-Manin connection as the connection arising by restricting a certain
translation-invariant connection to one specific $K$-point of the generalized
jacobian. The essential point seems to be that the de Rham cohomology of a
connection of the form $d+\omega$ on a trivial bundle is somehow concentrated
at the points where $\omega = 0$.
In section 4 we reinterpret the Riemann-Roch formula in terms of a pairing
\eqref{4.5}
\begin{gather}
\cup: \H^1(C, {\mathcal O}^*_C \to {\mathcal O}^*_{{\mathcal D}}) \times
\H^1(C, {\mathcal O}^*_C \to \Omega^1_C\langle D \rangle({\mathcal D}')) \notag \\
\to \H^2(C, {\mathcal K}_2 \to \Omega^2_C)
\end{gather}
and a trace map \eqref{4.6}
$${\rm Tr \ }:\H^2(C, {\mathcal K}_2 \to \Omega^2_C) \to \Omega^1_K/d\log K^*.
$$
In section 5 we give an analogous ``non-commutative''
product formula in the higher rank case which
we conjecture calculates the determinant connection in the generic situation
when the connection defines local isomorphisms $E|_{\mathcal D} \cong E|_{\mathcal D}\otimes
\omega_{{\mathcal D}/K}$ (see \eqref{ass5.3}) and the poles of the
absolute connection behave well (see \eqref{ass5.1}).
We verify the formula has the appropriate invariance properties.
We also show that there is a more general higher rank product of
which it is a special case.
Finally, in section 6 we give a general formula which calculates the group of
isomorphism classes of irregular, integrable, rank $1$ connections in higher
dimensions on a smooth projective variety.
We apologize for not expressing our results in the modern language of
${\mathcal D}$-modules, but in fact for the study of Gau{\ss}-Manin determinants
there is little gain in passing from connections to ${\mathcal D}$-modules. Also,
rigidity for connections means that the Gau{\ss}-Manin
determinant connection is
determined by its value at the generic point on the base, so we may work
with curves over a function field ${\rm Spec \,}(K)$.
It is our pleasure to
acknowledge the intellectual debt we owe in this work to P. Deligne. We
are also grateful to the Humboldt foundation for financing which enabled
us to work together.
\section{Connections and forms on Generalized Jacobians}
Throughout this paper $C$ will be a smooth projective curve over a field
$K$ containing an algebraically closed subfield $k$ of
characteristic 0, and ${\mathcal D} = \sum m_i c_i$ is a divisor on $C$,
with $c_i \in C(K)$.
We write $G=J_{{\mathcal D}}$ for the
generalized Jacobian parametrizing isomorphism classes of degree $0$
line bundles on $C$ with trivialization along ${\mathcal D}$. Fixing a
$K$-rational point $$c_0 \in (C-D)(K),$$ there is
a cycle map $i: C-D \to J_{{\mathcal D}}$ associating to a closed point $x\in C$ with
$[K(x):K]=n$ the class of the line bundle ${\mathcal O}(x-nc_0)$ together with the
trivialization $b|_{\mathcal D}\circ(a|_{\mathcal D})^{-1}$, where
$${\mathcal O}_C \stackrel{a}{\hookleftarrow} {\mathcal O}_C(-nc_0) \stackrel{b}{\hookrightarrow}
{\mathcal O}_C(x-nc_0)$$
are the natural maps.
The aim of this section is to describe invariant line bundles
with connection on $J_{{\mathcal D}}$, comparing them via the cycle
map $i$ to line bundles with connection on $(C-D)$ with a
certain irregularity behavior along $D$.
When the line bundle in question is the trivial bundle, this amounts to
studying invariant (absolute) differential forms on the generalized
jacobian, so we should start with that. Before doing so, however, it is
necessary to understand global functions on the generalized
jacobian. We write
\eq{G \twoheadrightarrow G_0 \twoheadrightarrow J}{2.1}
where $J$ is the usual Jacobian of $C$, and $G_0$ is a semi-abelian
variety. We have extensions
\begin{gather}
0 \to T \to G_0 \to J \to 0 \label{2.2}\\
0 \to {\mathbb V} \to G \to G_0 \to 0 \label{2.3}
\end{gather}
Here ${\mathbb V}$ is a vector group (isomorphic to ${\rm Spec \,}(\text{Sym}(V^*))$ for some
vector space $V$) and $T$ is a torus, i.e. $T_{\bar{K}}\cong
{\mathbb G}_m^r$.
\begin{lem}\label{lem2.1}The semi-abelian variety $G_0$ admits a universal
vectorial extension
\eq{0 \to {\mathbb W} \to {\mathcal G} \to G_0 \to 0.
}{2.4}
In fact, this extension is given by the pullback to $G_0$ of the
universal vectorial extension over $J$. In particular, ${\mathbb W} =
\Gamma(J^\vee,\Omega^1_{J^\vee/K})\otimes {\mathbb G}_a$.
\end{lem}
\begin{proof} It will suffice to show the pullback vectorial extension
is universal. Since $\text{Ext}^1(T,{\mathbb G}_a)=(0)=\text{Hom}(T,{\mathbb G}_a)$, any
extension of $G_0$ by a vector group ${\mathbb W}$ is pulled back from a unique
extension of $J$ by ${\mathbb W}$. This extension of $J$ is a pushout from the
universal vectorial extension, so the same holds for the pullbacks to
$G_0$.
\end{proof}
\begin{lem}\label{lem2.2} Let $\pi:G_0\to J$ be an extension of $J$ by
$T$ as above. There
exists, possibly after a finite field extension,
a quotient torus $T \twoheadrightarrow S$ and a diagram
\eq{\begin{array}{ccc} \makebox[0cm][c]{$T$} & \hookrightarrow &
\makebox[0mm]{$G_0$} \\
\ \ \ \makebox[0mm][r]{surj. $\downarrow$ } & \swarrow a \\
\makebox[0cm][c]{$S$}
\end{array}
}{2.5}
such that
\eq{H^i(G_0,{\mathcal O}_{G_0})\cong H^i(J,{\mathcal O}_J) \otimes_{K} H^0(S,{\mathcal O}_S)
}{2.6}
\end{lem}
\begin{proof}There is a boundary map
\eq{\partial:{\rm Hom}_{\bar{K}}(T,{\mathbb G}_m) \to
{\rm Ext}^1_{\bar{K}}(J, {\mathbb G}_m)
}{2.7}
Define $N:=\ker(\partial) \subset M:=Hom_{\bar{K}}(T,{\mathbb G}_m) $.
Let $S= \text{Hom}(N,{\mathbb G}_m)$ be the torus with character group
$N$.
For $m\in M$ let $L(m)$ be the line bundle on $J_{\bar{K}}$
corresponding under the map \eqref{2.7}. As an ${\mathcal O}_{J_{\bar K}}$-algebra
\eq{\pi_*{\mathcal O}_{G_{0,\bar K}} \cong \oplus_{m\in M}L(m) \cong
H^0({\mathcal O}_S)\otimes \Big(\oplus_{m\in M/N}L(m)\Big)
}{2.8}
The map $a$ in the diagram \eqref{2.5} comes from the above inclusion
$$H^0({\mathcal O}_S)\otimes L(0) \subset \pi_*{\mathcal O}_{G_0}.$$
For $m\in M/N$, (as is well known, cf. \cite{Mu} III 16),
$L(m)$ has trivial cohomology in all
degrees unless $m=0$. The proposition follows by taking cohomology of
\eqref{2.8}.
\end{proof}
\begin{lem}\label{lem2.3} Let notation be as above. Let
\eq{
\begin{CD}0 @>>> H^0(J^\vee,\Omega^1_{J^\vee/K})\otimes {\mathbb G}_a @>>>
{\mathcal G} @>p>> G_0 @>>> 0\end{CD}
}{2.9}
be the universal vectorial extension. Then
\eq{
H^0({\mathcal G},{\mathcal O}_{\mathcal G})\cong H^0(G_0, {\mathcal O}_{G_0})
}{2.10}
\end{lem}
\begin{proof} The ${\mathcal O}_{G_0}$-algebra $p_*{\mathcal O}_{\mathcal G}$ is filtered, with
\eq{
gr_ip_*{\mathcal O}_{\mathcal G} = fil_i/fil_{i-1} \cong
\text{Sym}^i(H^0(J^\vee,\Omega^1_{J^\vee/K})^*)\otimes {\mathcal O}_{G_0}
}{2.11}
With respect to the exact sequences
\eq{\begin{CD} 0 @>>> fil_{i-1} @>>> fil_i @>>> gr_i @>>> 0 \end{CD}
}{2.12}
it suffices to show the boundary map
\eq{b: \text{Sym}^i(H^0(J^\vee,\Omega^1_{J^\vee/K})^*)\otimes H^0(G_0,{\mathcal O}_{G_0}) \to
H^1(G_0, fil_{i-1})
}{2.13}
is injective. Composing on the right with the evident map, it suffices
to show the maps
\ml{\text{Sym}^i(H^0(J^\vee,\Omega^1_{J^\vee/K})^*)\otimes H^0(G_0,{\mathcal O}_{G_0}) \\
\to
H^1(G_0, {\mathcal O}_{G_0})\otimes
\text{Sym}^{i-1}(H^0(J^\vee,\Omega^1_{J^\vee/K})^*)
}{2.14}
are injective. But
\ml{H^0(J^\vee,\Omega^1_{J^\vee/K})^*\otimes H^0(G_0,{\mathcal O}_{G_0}) \cong
H^1(J, {\mathcal O}_J)\otimes H^0(G_0,{\mathcal O}_{G_0}) \\
\cong H^1(G_0, {\mathcal O}_{G_0})
}{2.15}
and the map in \eqref{2.14} is the map $x^i\mapsto x\otimes x^{i-1}$,
which is injective.
\end{proof}
\begin{lem}\label{lem2.4} Let $G=J_{\mathcal D}$ be a generalized jacobian as
above. Then there exists a commutative affine algebraic group ${\mathbb G}$ over
$K$ and a map $\psi : G \to {\mathbb G}$ such that
\eq{\psi^* : H^0({\mathcal O}_{\mathbb G}) \cong H^0({\mathcal O}_G).
}{2.16}
\end{lem}
\begin{proof}Take ${\mathbb G} = {\rm Spec \,}(H^0(G,{\mathcal O}_G))$.
\end{proof}
\begin{lem}\label{lem2.5} Let $A$ be the coordinate ring of a
commutative affine algebraic group $H$ over a field $K$ of
characteristic $0$. Corresponding to the simplicial algebraic group
$BH$, one has a complex
\eq{
\begin{CD} A @>\mu^*-p^*_1 -p^*_2 >> A\otimes_K A @> p_{23}^* -
\mu_{12}\otimes p_3^* + p_1^*\otimes \mu_{23}^* - p_{12}^* >> A\otimes
A\otimes A \end{CD}
}{2.19}
This complex is exact at the middle term.
\end{lem}
\begin{proof} By proposition 4 on p. 168 of \cite{Se},
the cohomology in the middle is a
subgroup of the group of extensions $Ext(H,{\mathbb G}_a)$. (Note, $A
=Map(H,{\mathbb G}_a)$.) By the classification of commutative algebraic groups
in characteristic $0$, this ext group vanishes (cf \cite{Se}, pp. 170-172).
\end{proof}
We write $\Omega^1_G$ (resp. $\Omega^1_{G/K}$) for the sheaf of $1$-forms relative to
$k$ (resp. $K$).
Now we would like to define invariant bundles, connections, differential forms,
cohomology classes of ${\mathcal O}_G$.
\begin{defn}\label{def2.6}\begin{enumerate}
\item A rank one bundle $L \in H^1(G, {\mathcal O}^*_G)$ is called invariant
if $\mu^*L= p_1^*L \otimes p_2^*L \in H^1(G\times G, {\mathcal O}^*_{G
\times G})$, where $\mu: G \times G \to G$ is the multiplication
and $p_i: G \times G \to G$ are the projections.
\item A global $1$-form $\eta\in \Gamma(G, \Omega^1_G)$ is called
invariant if
$\eta(0)=0 \in \Omega^1_K$ and $\mu^*\eta = p_1^*\eta + p_2^* \eta \in
\Gamma(G\times G, \Omega^1_{G\times G})$.
\item A rank one bundle with a connection $(L, \nabla) \in \H^1(G,
{\mathcal O}^*_G \to \Omega^1_G)$ is called invariant if $(L,
\nabla)|\{0\}= 0 \in \H^1({\rm Spec \,} K, {\mathcal O}^*_{{\rm Spec \,} K} \to
\Omega^1_{{\rm Spec \,} K}) = \Omega^1_K /d\log K^*$ and
$\mu^*(L, \nabla)= p_1^*(L, \nabla) \otimes p_2^*(L, \nabla)
\in \H^1(G \times G, {\mathcal O}^*_{G \times G} \to \Omega^1_{G \times
G})$.
\item A class $s\in H^i(G, {\mathcal O}_G)$ is called
invariant if $s|\{0\}=0$ and
$\mu^* s= p_1^*s + p_2^*s$ in $H^i(G\times G, {\mathcal O}_{G \times
G})$.
\end{enumerate}
We denote by $H^1(G, {\mathcal O}^*_G)^{{\rm inv}}$,
$\Gamma(G, \Omega^1_G)^{{\rm inv}}$, $\H^1(G,
{\mathcal O}^*_G \to \Omega^1_G)^{{\rm inv}}$, and $H^i(G, {\mathcal O}_G)^{{\rm inv}}$
the corresponding groups of
invariant bundles, forms, connections and classes.
One defines similarly the
groups of relative invariant forms
$H^0(G, \Omega^1_{G/K})^{{\rm inv}}$ and relative invariant
connections $\H^1(G, {\mathcal O}^*_G \to \Omega^1_{G/K})^{{\rm inv}}$
without condition on the restriction to the zero section, and
observe that the natural map
$\Omega^1_G \to \Omega^1_{G/K}$ takes global invariant groups to
relative invariant groups.
\end{defn}
\begin{remark} In the above definitions, we could have defined a weaker notion
of invariance by allowing constant elements.
We adopt here the rigidification at the origin, keeping in mind that
without this condition, the corresponding groups obtained
are a direct sum of the ones obtained with the rigidification
and the value of the group on the zero section. Notice, for example, that with
our definition, nonzero constant functions are not invariant! Of course, for
relative objects, there is no distinction.
\end{remark}
\begin{lem}\label{lem2.7} Let $G=J_{{\mathcal D}}$ be a generalized Jacobian as
above. Let
$\tau \in \Gamma(G, \Omega^1_{G/K})^{{\rm inv}}$ be an invariant relative $1$-form on $G$, and
assume $\tau $ lifts to an absolute global form. Then $\tau$ lifts to
an invariant absolute form on $G$.
\end{lem}
\begin{proof}Let $\eta \in \Gamma(G, \Omega^1_G)$ be an absolute lifting. Replacing
$\eta$ with $\eta - \eta(0)$ we may assume $\eta(0)\in \Omega^1_K$
vanishes. Then
\eq{(\mu^*-p_1^* -p_2^*)(\eta)\in H^0({\mathcal O}_{G\times_K G})\otimes\Omega^1_K
}{2.20}
vanishes in $H^0({\mathcal O}_{G\times_K G\times_K G})\otimes\Omega^1_K$. Let $\psi: G
\to {\mathbb G}$ be as in lemma \ref{lem2.4},
so $\psi^*:A := H^0({\mathcal O}_{{\mathbb G}}) \cong H^0({\mathcal O}_G)$. The
previous lemma implies there exists $\sigma \in H^0({\mathcal O}_G)\otimes \Omega^1_K$
with $(\mu^*-p_1^* -p_2^*)(\sigma) = (\mu^*-p_1^* -p_2^*)(\eta)$. Then
$\eta - \sigma$ is the desired invariant absolute form.
\end{proof}
We next need to relate connections on the curve with invariant connections
on the generalized Jacobian. Here $G = J_{{\mathcal D}}$ with ${\mathcal D} = \sum
m_i c_i$. Also,
\eq{D := \sum c_i;\quad {\mathcal D}' := {\mathcal D} -D.
}{2.21}
We assume $D\neq \emptyset$.
First, rather briefly, we consider the question of what poles
invariant forms on $G$ have when pulled back to $C-D$ via the jacobian
map $C-D \to G$ (defined once we have a basepoint in $C-D$). For simplicity, we
continue to assume the $c_i$ are defined over $K$. Consider the diagram:
\eq{\begin{array}{ccccc} G \twoheadrightarrow G_0
& \twoheadrightarrow & J \\
\makebox[1.3cm][l]{$\uparrow i$} & & \makebox[.2cm][l]{$\uparrow i'$} \\
\makebox[0cm][c]{$C-D\quad $} & \stackrel{j}{\hookrightarrow} & C
\end{array}
}{2.22}
We want to compute the pullbacks $i^*H^0(\Omega^1_G)^{inv}$ (resp.
$i^*H^0(\Omega^1_{G/K})^{inv}$) in $H^0(C,\Omega^1_C(*D))$ (resp. in
$H^0(C,\Omega^1_{C/K}(*D))$) .
Pulling back $G$ via $i'$ we get a torseur
\eq{{i'}^*G \stackrel{p}{\to} C
}{2.23}
under the group ${\mathcal O}^*_{{\mathcal D}}/{\mathbb G}_m = \prod {\mathcal O}^*_{m_\ell c_\ell}/{\mathbb G}_m$. Fix
$\ell$ and let $R=k[[t_\ell]]\subset M=k((t_\ell))$ where $t_\ell$ is a formal
parameter at
$c_\ell$. Fix a splitting of the torseur
\eq{G_R \cong {\mathcal O}^*_{{\mathcal D}}/{\mathbb G}_m \times {\rm Spec \,}(R).
}{2.24}
Let $c_0\in (C-D)(K)$ be the base point used to define the
map $i:C-D \to G$. Let ${\mathcal L}$ on $C\times C$ be a line bundle with
${\mathcal L}|_{\{c\}\times C}\cong {\mathcal O}(c-c_0)$. Note one has
\eq{{\mathcal O}_{C\times C}\leftarrow p_2^*{\mathcal O}(-c_0) \to {\mathcal L}
}{2.25}
and these maps are isomorphisms on
\eq{{\rm Spec \,}(M)\times {\rm Spec \,}({\mathcal O}_{C,D}) \subset
{\rm Spec \,}(M)\times C
}{2.26}
Corresponding to ${\mathcal L}|_{{\rm Spec \,}(M)\times C}$ and the
above trivialization, one gets a map $u: {\rm Spec \,}(M) \to G$. With respect
to the above splitting, we view $u$ as an element
\eq{\prod_i (u_{i0}+u_{i1}t_i+\ldots+u_{i,m_i-1}t_i^{m_i-1})\in
\prod_i ({\mathcal O}_{m_ic_i}\otimes_K M)^* \mod M^*
}{2.27}
As described in \cite{Se} VII 4, 21, the local shape of $u$
around $c_\ell$ is given by taking the rational function
$(s_\ell-t_\ell)^{-1} $, where the local coordinates around
$(c_\ell, c_\ell)$ in $C \times C$ is $(s_\ell, t_\ell)$, and
considering it as a unit in ${\mathcal O}_{m_\ell c_\ell}\otimes M
\cong M[t_\ell]/<t_\ell^{m_\ell}>$. (We change notation so $s_\ell$ is the
local parameter in $R\subset M$.) Since
$u$ is well defined and non-vanishing in $c_i$ for $i\neq \ell$, we have
\eqref{2.26} that
\eq{u_{ij}\in R, u_{i0}\in R^*\text{ if } i\neq \ell,\
\text{ord} (u_{\ell
0})=1}{2.28}
The pullbacks to ${\rm Spec \,}(M)$ of the invariant relative differential
forms on $G$ are given by the pullback of invariant relative forms on
$J$ together with the coefficients of powers of the $T_i \mod
T_i^{m_i} $ in the formal
expression
\begin{multline}\label{2.29}
\sum_i
(u_{i0}+u_{i1}T_i+\ldots+u_{i,m_i-1}T_i^{m_i-1})^{-1} \times \\
(du_{i0}+ du_{i1}T_i+
\ldots+
du_{i,m_i-1}T_i^{m_i-1}) \\
= \sum_i \tau_{i0}+\ldots + \tau_{i,m_i-1}T_i^{m_i-1}.
\end{multline}
Then \eqref{2.28} implies that
\begin{gather}\label{2.30}
\tau_{ij} \in \Omega^1_{R} \ {\rm for \ } i \neq \ell \notag \\
\tau_{\ell j} \in \Omega^1_R\langle D \rangle( jD_\ell)-\Omega^1_R\langle D \rangle( (j-1)D_\ell).
\end{gather}
Here we denote by
$\Omega^1_C\langle D \rangle$ the sheaf of absolute differential forms of degree
1 with logarithmic poles along $D$. (See formula \ref{pbtau} for a more
precise computation).
These are not all the
absolutely invariant forms, however. One also has forms pulled back
from $J$, but these are regular along $D$. Finally, from lemma
\ref{lem2.7} one has an exact sequence
\ml{0 \to H^0({\mathcal O}_G)^{{\rm inv}} \otimes \Omega^1_K \to
H^0(G,\Omega^1_G)^{{\rm inv}}
\to H^0(G,\Omega^1_{G/K})^{{\rm inv}} \\
\to H^1({\mathcal O}_G)^{{\rm inv}} \otimes \Omega^1_K
}{2.31}
It shows one must consider invariant forms in
$H^0({\mathcal O}_G)^{{\rm inv}}\otimes \Omega^1_K$. We will see in the proof of
proposition \ref{prop2.12} below that these map to $\Omega^1_G({\mathcal D}')$. In sum,
the above discussion shows that the maps in the following proposition
are defined.
\begin{prop}Pullback gives isomorphisms
\begin{gather}\label{2.32}H^0(\Omega^1_{G/K})^{inv}
\stackrel{\cong}{\longrightarrow} H^0(C, \Omega^1_{C/K}({\mathcal D})) \\
\label{2.33} H^0(\Omega^1_G)^{inv} \stackrel{\cong}{\longrightarrow} H^0(C,
\Omega^1_C\langle D \rangle({\mathcal D}'))
\end{gather}
\end{prop}
\begin{proof} Pullback on invariant relative forms is injective,
because $G$ is
generated by the image of $C-D$. It follows by dimension count that
the first arrow \eqref{2.32}
above is an isomorphism. For the absolute forms we may
consider the diagram
\begin{tiny}
\begin{equation}\label{2.34}\begin{array}{ccccccccc}
0 & \to & H^0({\mathcal O}_G)^{{\rm inv}}\otimes\Omega^1_K & \to & H^0(\Omega^1_G)^{{\rm inv}} &
\makebox[0cm][r]{$\to$} & H^0(\Omega^1_{G/K})^{{\rm inv}} & \to & \quad\quad\ \
\makebox[1.3cm][r]{$H^1({\mathcal O}_G)^{{\rm inv}}\otimes
\Omega^1_K$} \\
&& \downarrow \cong && \downarrow && \downarrow \cong &&
\downarrow
\cong \\
0 & \to & \makebox[1cm][c]{$H^0({\mathcal O}_C({\mathcal D}'))\otimes\Omega^1_K$} & \to &
\makebox[2.2cm][c]{$\ \ H^0(\Omega^1_C\langle D \rangle({\mathcal D}'))$} & \makebox[0cm][r]{$\to$} &
H^0(\Omega^1_{C/K}({\mathcal D})) &\ \makebox[0cm][r]{$\to$} & \quad\quad\ \
\makebox[1.3cm][r]{\ \ $H^1({\mathcal O}_C({\mathcal D}'))\otimes\Omega^1_K$}
\end{array}
\end{equation}
\end{tiny}
The left and right hand vertical arrows are shown to be isomorphisms
in the proof of proposition \ref{prop2.12}. Hence the isomorphism on invariant
relative forms implies the isomorphism \eqref{2.33}
on invariant absolute forms.
\end{proof}
We now consider invariant connections on line bundles on $G$.
\begin{lem}\label{lem2.8}Assume the toric subquotient $T$ of $G$ has
trivial Picard group (e.g. $T$ split). Then the map $\H^1(G, {\mathcal O}_G^*
\to \Omega^1_{G/K})^{{\rm inv}} \to H^1({\mathcal O}_G^*)^{{\rm inv}}$ is surjective.
\end{lem}
\begin{proof}This follows because
\eq{\H^1(J, {\mathcal O}_J^* \to \Omega^1_{J/K})^{{\rm inv}} \twoheadrightarrow
H^1({\mathcal O}_J^*)^{{\rm inv}}\twoheadrightarrow H^1({\mathcal O}_G^*)^{{\rm inv}}.
}{2.35}
The second arrow is surjective because we have a diagram
\eq{\begin{CD}0 @>>> H^1({\mathcal O}_J^*)^{{\rm inv}} @>>> H^1({\mathcal O}_J^*) @>>>
H^2_{DR}(J/K) \\
@. @VVV @VV surj. V @VV inj. V \\
0 @>>> H^1({\mathcal O}_G^*)^{{\rm inv}} @>>a > H^1({\mathcal O}_G^*) @>>b >
H^2_{DR}(G/K) \\
\end{CD}
}{2.36}
The bottom row is not a priori exact, but $b\circ a=0$ (because
$(N\delta)^*$ acts by $N^2$ on $H^2_{DR}(G/K)$.) The middle
vertical arrow is onto e.g. because the Picard group of the generic
fibre of $G \to J$ is zero. Indeed, $G\to J$ is rationally split, and
the kernel has trivial Picard group by hypothesis. (Since the function
field of the generic fibre equals the function field of $G$, any
divisor on $G$ can be moved by rational equivalence to avoid the
generic fibre, i.e. to be a pullback from the base.)
Finally, the right hand vertical
arrow is injective because, after making a base change
$K \subset {\mathbb C}$, one can think of $G$ and $J$ as quotients
of vector spaces by lattices, and the map on lattices is surjective
$\otimes {\mathbb Q}$.
\end{proof}
\begin{lem}\label{lem2.10} Let $a\in \H^1(G, {\mathcal O}^*_G \to \Omega^1_{G/K})^{{\rm inv}}$
be an invariant connection on a line bundle on the generalized jacobian
$G$. Suppose $a$ lifts to an absolute connection $b'\in \H^1(G,
{\mathcal O}^*_G \to \Omega^1_G)$. Then $a$ lifts to an invariant absolute connection
$b\in \H^1(G, {\mathcal O}^*_G \to \Omega^1_G)^{{\rm inv}}$.
\end{lem}
\begin{proof} One has as in \eqref{2.19}
\ml{(\mu^*-p_1^*-p_2^*)(b')\in \
{\rm Im \ }H^0({\mathcal O}_{G\times G})\otimes \Omega^1_K \\
\subset
\H^1(G\times G, {\mathcal O}^*_{G\times G} \to \Omega^1_{G\times G})
}{2.37}
Now
\begin{gather}
{\rm Im \ }H^0({\mathcal O}_{G\times G})\otimes \Omega^1_K = \notag \\
H^0({\mathcal O}_{G\times G})\otimes \Omega^1_K/d\log {\rm Ker \ }
\{H^0({\mathcal O}_{G\times G}^*) \to H^0(\Omega^1_{G \times G/K})\}= \notag \\
H^0({\mathcal O}_{G\times G})\otimes (\Omega^1_K/d\log K^*). \notag
\end{gather}
Exactness of the sequence in \eqref{2.19} implies that
there exists an element $x \in H^0(G, {\mathcal O}\otimes \Omega^1_K)$ with
$(\mu^*-p_1^*-p_2^*)(x)=(\mu^*-p_1^*-p_2^*)(b')$. Take $b=b'-x$.
\end{proof}
\begin{prop}\label{prop3} One has an exact sequence
\begin{multline}\label{2.38} 0 \to H^0({\mathcal O}_G)^{{\rm inv}}\otimes \Omega^1_K \to
\H^1(G,{\mathcal O}_G^* \to \Omega^1_G)^{{\rm inv}} \\
\to \H^1(G,{\mathcal O}_G^* \to \Omega^1_{G/K})^{{\rm inv}} \to H^1({\mathcal O}_G)^{{\rm inv}}\otimes \Omega^1_K
\end{multline}
\end{prop}
\begin{proof} This is immediate from the lemma.
\end{proof}
Recall our notation. $C$ is a smooth, projective, geometrically
connected curve over $K$. $G=J_{\mathcal D}$ with ${\mathcal D}=\sum m_i c_i$, $D=\sum c_i$,
${\mathcal D}' = {\mathcal D} - D$.
\begin{prop}\label{prop2.12} There exists a diagram of exact sequences,
with vertical
arrows isomorphisms:
\begin{tiny}
\eq{\begin{array}{cccccc}
0 & \to & H^0({\mathcal O}_G)^{{\rm inv}} \otimes (\Omega^1_K/d \log K^*)
& \to & \H^1(G,{\mathcal O}_G^* \to
\Omega^1_G)^{{\rm inv}} \\
&& \rule{0cm}{.5cm}\downarrow e && \downarrow \\
0 & \to & \Big(H^0({\mathcal O}_C({\mathcal D}'))/H^0({\mathcal O}_C)\Big) \otimes (\Omega^1_K/d \log K^*) & \to &
\rule{0cm}{.5cm}\makebox[3.5cm][c]{$\frac{\H^1(C,j_*({\mathcal O}_{C-D}^*) \to
\Omega^1_C\langle D \rangle({\mathcal D}'))}{(\Omega^1_K/K^\times)}$} \\
\rule{0cm}{1cm} \\
& \to & \H^1(G,{\mathcal O}_G^* \to \Omega^1_{G/K})^{{\rm inv}} & \to &
H^1({\mathcal O}_G)^{{\rm inv}}\otimes \Omega^1_K \\
&& \rule{0cm}{.5cm}\downarrow && \downarrow h\\
&\to & \rule{0cm}{.5cm}\H^1\Big(C,j_*({\mathcal O}_{C-D}^*)\to
\Omega^1_{C/K}({\mathcal D})\Big) & \to &
H^1({\mathcal O}_C({\mathcal D}'))\otimes \Omega^1_K \\
\end{array}
}{2.39}
\end{tiny}
\end{prop}
\begin{proof} The first step is to compute $H^i({\mathcal O}_G)^{{\rm inv}}$ for
$i=0,1$. Let $W$ be a finite dimensional
$K$-vector space, and suppose $G=J_{\mathcal D}$ is a vectorial extension
\eq{0 \to W\otimes {\mathbb G}_a \to G \stackrel{p}{\to} G_0 \to 0
}{2.40}
We know by lemma \ref{lem2.1} that this sequence pulls back from
an extension of $J$ by $W\otimes {\mathbb G}_a$. Let
\eq{0 \to {\mathcal O}_J \to fil_1 \to W^*\otimes {\mathcal O}_J \to 0
}{2.41}
be the exact sequence of functions of filtration degree $\le 1$ as in
lemma \ref{lem2.3}, and let $\partial : W^* \to H^1({\mathcal O}_J)$ be the
boundary map in cohomology.
\begin{lem}We have
\begin{gather}H^0({\mathcal O}_G)^{{\rm inv}} \cong
\label{2.42}\ker(\partial: W^* \to H^1({\mathcal O}_J)) \\
\label{2.43} H^1({\mathcal O}_G)^{{\rm inv}} \cong H^1({\mathcal O}_J)/\partial(W^*).
\end{gather}
\end{lem}
\begin{proof}[proof of lemma]
One has a filtration $fil_\cdot p_*{\mathcal O}_G$ with $fil_0 = {\mathcal O}_{G_0}$ and
$gr_r = \text{Sym}^r(W^*)\otimes {\mathcal O}_{G_0}$. The corresponding spectral
sequence looks like
\begin{equation}\label{2.44} E_1^{pq} = H^{p+q}(G_0, gr_{-p}) =
H^{p+q}(G_0, \text{Sym}^{-p}(W^*)\otimes {\mathcal O}_{G_0}) \Rightarrow
H^{p+q}({\mathcal O}_G).
\end{equation}
Let
\eq{0 \to H_0 \to G_0 \to S \to 0
}{2.45}
be as in lemma \ref{lem2.2}, so $S$ is the maximal quotient torus
of $G_0$.
The equation \eqref{2.6} identifies
$H^i(G_0, {\mathcal O}_{G_0})$ with $H^i(J \times S, {\mathcal O}_{J \times S})$,
and the invariance condition might be looked at on $J \times S$.
Let us write $H^0(S, {\mathcal O}_S) = K \oplus V, f \mapsto f(0) \oplus (f - f(0))$,
where $V$ consists of the regular functions which vanish
at $\{0\} \in S$.
Then $H^i(G_0, {\mathcal O}_{G_0})^{{\rm inv}}= H^i(J, {\mathcal O}_J)^{{\rm inv}}
\oplus (H^i(J, {\mathcal O}_J) \otimes V)^{{\rm inv}}$.
Thus if a class $F= \sum \varphi_f \otimes f \in H^i({\mathcal O}_J)
\otimes V$ is invariant, where $\varphi_f \in H^i(J, {\mathcal O}_J)$
and the $f \in V$ are linearly independent over $K$,
then
\begin{gather}
(\mu^* - p_1^* -p_2^*)(F)|(J\times J \times \{0\} \times S)
= \notag \\
\sum (\mu^* - p_2^*)(\varphi_f) \otimes f = 0 \notag
\end{gather}
thus $\mu^*\varphi_f=p_i^*\varphi_f$ and
$\mu^*\varphi_f|\{0\}\times J= \varphi_f = \varphi_f|\{0\}.$
So for $i\geq 1$, this implies that $\varphi_f=0$ and for $i=0$
this implies that $F\in H^0({\mathcal O}_S)^{{\rm inv}}=0.$
In short:
\eq{H^*({\mathcal O}_{G_0})^{{\rm inv}} \cong \Big(H^*({\mathcal O}_J)\otimes
H^0({\mathcal O}_S)\Big)^{{\rm inv}} \cong H^*({\mathcal O}_{H_0})^{{\rm inv}}
\cong H^*({\mathcal O}_J)^{{\rm inv}}
}{2.46}
Thus, it suffices to prove the lemma with $G_0$ replaced by
$H_0$, so we may assume the quotient torus $S=(0)$.
Since in this case $H^i({\mathcal O}_{G_0})\cong \wedge^i H^1({\mathcal O}_J)$, one sees
that
pullback under the multiplication by $N$ map, $N\delta:G \to G$ acts
on $E_1^{pq}$ by multiplication by $N^q$. It follows that the
spectral sequence \eqref{2.44} degenerates at $E_2$. In particular the
eigenspace where $N\delta^*$
acts by multiplication by $N$ on $H^1({\mathcal O}_G)$ is
\begin{equation}\label{2.47} H^1({\mathcal O}_J)/\partial(W^*) \cong E_2^{0,1}
\cong E_\infty^{0,1} \hookrightarrow H^1({\mathcal O}_G).
\end{equation}
Note that as a quotient of $H^1({\mathcal O}_J)$ the space $E_\infty^{0,1}$ is
clearly invariant. Conversely, let $\Delta: G \to G\times G$ be the
diagonal. Since $\mu\circ\Delta = 2\delta$ it follows that for $a\in
H^1({\mathcal O}_G)^{{\rm inv}}$ we have
\eq{(2\delta)^*(a) = \Delta^*\mu^*(a) = \Delta^*(p_1^*(a)+p_2^*(a)) = 2a
}{2.48}
so necessarily $a\in E_\infty^{0,1}$, proving \eqref{2.43}. A similar
argument on $E_\infty^{-1,1}$ proves \eqref{2.42}.
We remark here that $H^0(J, fil_1)$ is in a natural way a subspace of
the regular functions on $G$, and \eqref{2.42} takes the quotient
of this by $H^0(J, {\mathcal O}_J)= K$. This is because we have forced the rigidification
condition in the definition \ref{def2.6}.
\end{proof}
We return to the proof of proposition \ref{prop2.12}. The exact
sequence
\begin{equation}\label{2.49}
0 \to {\mathcal O}_C \to {\mathcal O}_C({\mathcal D}') \to {\mathcal O}_C({\mathcal D}')/{\mathcal O}_C \to 0
\end{equation}
defines a map
\begin{equation}\label{2.50}
\psi: W^* := H^0({\mathcal O}_C({\mathcal D}')/{\mathcal O}_C ) \to H^1({\mathcal O}_C).
\end{equation}
By lemma \ref{lem2.1}, as a group extension of $G_0$,
the group $G$ is defined by a unique map
from $H^0(C, \Omega^1_{C/K})$ to a vector space. We claim that
this map is the dual of $\psi$. To see this, one identifies $J$
and $J^\vee$. Then it is well know that the universal
vectorextension on $J^\vee$ is
$$0 \to H^0(C, \Omega^1_{C/K}) \to \H^1(C, {\mathcal O}^*_C \to
\Omega^1_{C/K}) \to {\rm Pic}^0(C) \to 0$$
inducing the universal vectorextension
$$0 \to H^0(C, \Omega^1_{C/K}) \to \H^1(C, {\mathcal O}^*_{C,D} \to
\Omega^1_{C/K}) \to {\rm Pic}^0(C, D) \to 0$$
on
$${\rm Pic}^0(C, D):= {\rm Ker}\Big(H^1(C, {\mathcal O}^*_{C,D})
\to H^1(C, \Omega^1_{C/K})\Big)$$
where $${\mathcal O}^*_{C, Z}:= {\rm Ker} ({\mathcal O}^*_C \to {\mathcal O}^*_{Z})$$
for any subscheme $Z \subset C$.
The map of complexes
$$a: \{{\mathcal O}^*_{C, {\mathcal D}} \to 0\} \to \{ {\mathcal O}^*_{C, D} \to
\Omega^1_{C/K}|_{{\mathcal D}'}\}$$
induces an isomorphism on $\H^1$. Indeed, $a$ sends
the exact sequence
$$0 \to H^0(C, (1+{\mathcal O}_{{\mathcal D}'}(-D))) \to H^1(C, {\mathcal O}^*_{C, {\mathcal D}})
\to H^1(C, {\mathcal O}^*_{C,D}) \to 0$$
to the exact sequence
$$0 \to H^0(C, \Omega^1_{C/K}|_{{\mathcal D}'}) \to
\H^1(C, {\mathcal O}^*_{C, D} \to
\Omega^1_{C/K}|_{{\mathcal D}'}) \to H^1(C, {\mathcal O}^*_{C,D}) \to 0,$$
so one has just to see that
$$ d \log: H^0(C, (1+{\mathcal O}_{{\mathcal D}'}(-D))) \to
H^0(C, \Omega^1_{C/K}|_{{\mathcal D}'})$$
is an isomorphism.
But $ H^0(C, {\mathcal O}_{{\mathcal D}'}(-D)) \cong H^0(C, (1+{\mathcal O}_{{\mathcal D}'}(-D)))$ via
the exponential map and the quasiisomorphism
\cite{DeI}
$$\{{\mathcal O}_C(-{\mathcal D})\to \Omega^1_{C/K}(-{\mathcal D}')\} \to \{
{\mathcal O}_C(-D) \to \Omega^1_{C/K}\}$$
allows to conclude.
Define a bundle ${\mathcal E}$ on $C$ by pullback
\begin{equation}\label{2.51}\begin{CD}
0 @>>> {\mathcal O}_C @>>> {\mathcal E} @>>> W^*\otimes{\mathcal O}_C @>>> 0 \\
@. @| @VVV @VVV \\
0 @>>> {\mathcal O}_C @>>> {\mathcal O}({\mathcal D}') @>>> {\mathcal O}_C({\mathcal D}')/{\mathcal O}_C @>>> 0
\end{CD}
\end{equation}
Because of the isomorphism $H^1({\mathcal O}_J)\cong H^1({\mathcal O}_C)$, the top row
of the diagram \eqref{2.51} pulls back uniquely from an extension of
$W^*\otimes {\mathcal O}_J$ by ${\mathcal O}_J$. There is a unique vectorial extension
\begin{equation}\label{2.52}
0 \to W\otimes {\mathbb G}_a \to H \stackrel{r}{\to} J \to 0
\end{equation}
such that the above extension of vector bundles coincides with
\begin{equation}\label{2.53}
0 \to {\mathcal O}_J \to fil_1r_*{\mathcal O}_H \to W^*\otimes {\mathcal O}_J \to 0
\end{equation}
{From} this we get a diagram (defining $t$ and $u$. Here $i:C
\hookrightarrow J$)
\begin{equation}\label{2.54}
\begin{CD} 0 @>>> {\mathcal O}_J @>>> fil_1r_*{\mathcal O}_H @>>> W^* \otimes {\mathcal O}_J @>>>
0 \\
@. @VVV @VV t V @VV u V \\
0 @>>> {\mathcal O}_C @>>> {\mathcal O}_C({\mathcal D}') @>>> {\mathcal O}_C({\mathcal D}')/{\mathcal O}_C @>>> 0
\end{CD}
\end{equation}
We get a diagram with exact rows
\begin{tiny}
\begin{equation}\label{2.55}
\begin{array}{ccccccccccccc} 0 & \to & H^0({\mathcal O}_{J}) & \to &
H^0(fil_1r_*{\mathcal O}_H) & \to & H^0(W^*\otimes {\mathcal O}_{J}) &
\stackrel{\partial}{\to} & H^1({\mathcal O}_{J}) & \to \\
&&\downarrow \cong && \downarrow t && \downarrow \cong && \downarrow
\cong \\
0 & \to & H^0({\mathcal O}_C) & \to & H^0({\mathcal O}_C({\mathcal D}')) & \to &
H^0({\mathcal O}_C({\mathcal D}')/{\mathcal O}_C) & \to & H^1({\mathcal O}_C) & \to \\
\\
H^1({\mathcal O}_H)^{{\rm inv}} & \to & 0 \\
\downarrow v \\
H^1({\mathcal O}_C({\mathcal D}')) & \to & 0
\end{array}
\end{equation}
\end{tiny}
The diagram \eqref{2.55} gives isomorphisms
\begin{gather}\label{2.56}e: H^0({\mathcal O}_G)^{{\rm inv}} \cong \ker(\partial)
\cong {\rm Ker}(H^0({\mathcal O}_C({\mathcal D}')/{\mathcal O}_C)\to H^1({\mathcal O}_C)) \\
\label{2.57} h: H^1({\mathcal O}_G)^{{\rm inv}} \cong H^1({\mathcal O}_C({\mathcal D}')).
\end{gather}
These are two of the desired arrows for the diagram in the
proposition.
\begin{lem}The natural map on relative connections
\eq{\H^1(G, {\mathcal O}_G^* \to \Omega^1_{G/K})^{{\rm inv}} \to \H^1\Big(C,j_*({\mathcal O}_{C-D}^*) \to
\Omega^1_{C/K}({\mathcal D})\Big)
}{2.58}
is an isomorphism.
\end{lem}
\begin{proof}[proof of lemma] We note the following facts:
\begin{enumerate}\item $H^1(G, \Omega^1_{G/K})^{{\rm inv}}=\Big( H^0(G, \Omega^1_{G/K})^{{\rm inv}}\otimes
H^1({\mathcal O}_G) \Big)^{{\rm inv}}= (0)$.
\item $H^1({\mathcal O}_G^*)^{{\rm inv}}\cong H^1({\mathcal O}_{C-D}^*)$. Indeed, as remarked
in the proof of lemma \ref{lem2.8} one has $J(K) \twoheadrightarrow
H^1({\mathcal O}_G^*)^{{\rm inv}}$. One checks that the kernel is generated by
divisors of degree $0$ supported on $D$.
\item $H^0({\mathcal O}_G^*)/\text{consts.}\cong H^0({\mathcal O}_G^*)^{{\rm inv}}\cong
H^0({\mathcal O}_{C-D}^*)/\text{consts.}$.
\item
$$\Big(H^0(\Omega^1_{G/K})/d\log(H^0({\mathcal O}_G^*))\Big)^{{\rm
inv}}=H^0(\Omega^1_{G/K})^{{\rm inv}}/d\log
(H^0({\mathcal O}_G^*)^{{\rm inv}})
$$
(This is seen by noting $H^0({\mathcal O}_G^*)^{{\rm inv}}=
{\rm Hom}(G,{\mathbb G}_m)$, so one has
a homomorphism $\psi : G \to {\mathbb G}_m^r$ such that $\psi^*$ is an
isomorphism on global units modulo constants. The assertion then
reduces to the case
$G={\mathbb G}_m^r$, which is easy.)
\end{enumerate}
We build a diagram \begin{tiny}
\begin{equation}\label{2.59} \begin{array}{ccccccccc}
0 & \to & \frac{H^0(\Omega^1_{G/K})^{{\rm inv}}}{H^0({\mathcal O}_G^*)^{{\rm inv}}} & \to &
\H^1({\mathcal O}_G^* \to \Omega^1_{G/K})^{{\rm inv}} & \to & H^1({\mathcal O}_G^*)^{{\rm
inv}} & \to & 0 \\
&& \downarrow && \downarrow && \downarrow \\
0 & \to & \frac{H^0(\Omega^1_{C/K}({\mathcal D}))}{H^0({\mathcal O}_{C-D}^*)} & \to &
\H^1\Big(C,j_*({\mathcal O}_{C-D}^*) \to \Omega^1_{C/K}({\mathcal D})\Big) & \to &
H^1(C-D,{\mathcal O}^*) & \to & 0
\end{array}\end{equation}
\end{tiny}
Since the left and right hand arrows are isomorphisms, it follows that
the central arrow is as well, proving the lemma.
\end{proof}
The assertions of the proposition follow easily from the lemma.
\end{proof}
Finally, we need an analogous result for integrable connections. More
precisely, we consider a slightly weaker condition.
\begin{defn}\label{def2.15} Let $X$ be a variety over $K$. A
connection $\nabla : E \to E\otimes \Omega^1_X\otimes K(X)$ (so
possibly with poles) is said to have
vertical curvature if the curvature
\eq{\nabla^2 : E \to
E\otimes\Omega^2_X\otimes K(X)
}{2.60}
has values in the subsheaf $E\otimes\Omega^2_K\otimes K(X) \subset
E\otimes\Omega^2_X\otimes K(X)$.
The group of line bundles with vertical curvature will be denoted
$$\H^1(X,{\mathcal O}_X^* \to \Omega^1_X)^{{\rm vert \ }}$$
and similarly for invariant line bundles with vertical
curvature
$$\H^1(X,{\mathcal O}_G^* \to \Omega^1_G)^{{\rm inv,vert \ }}.$$
\end{defn}
\begin{prop}\label{prop2.16}
With notation as above, we have isomorphisms
\begin{gather}\label{2.61} \H^1(G, {\mathcal O}^*_G
\to \Omega^1_{G/K})^{{\rm inv,vert}}= \H^1(C, j_*{\mathcal O}^*_{C-D}
\to \Omega^1_{C/K}({\mathcal D}))^{{\rm vert}} \\
\label{2.62} \H^1(G, {\mathcal O}^*_G \to \Omega^1_G)^{{\rm inv}, {\rm vert \ }}\cong
\frac{\H^1(C,j_*{\mathcal O}_{C-D}^* \to
\Omega^1_C\langle D \rangle({\mathcal D}'))^{{\rm vert \ }}}{\Omega^1_K/K^\times}
\end{gather}
\end{prop}
\begin{proof}For example, in the absolute case, the curvature of a
line bundle with invariant absolute connection on $G$ is a
section $\eta \in H^0(\Omega^2_G/\Omega^2_K\otimes{\mathcal O}_G)$ satisfying
$\mu^*(\eta) = p^*_1(\eta)+p^*_2(\eta)$. It is easy to see that such a
section lies in the subsheaf $\Omega^1_{G/K}\otimes\Omega^1_K$. The isomorphism
\eqref{2.62} follows from proposition \ref{prop2.12} and the fact that
pullback to $C$ of invariant forms is injective by \eqref{2.32}. The
case of relative forms is similar and is left for the reader.
\end{proof}
\section{The Geometric Setup}
We continue to work with a curve $C/K$ and a line bundle $L$ on $C$
of degree $0$. Let $\nabla_{/K}:L \to L\otimes\Omega^1_{C/K}({\mathcal D})$, where ${\mathcal D} =
\sum m_ic_i$. As in the previous section, write $D=\sum c_i$ and ${\mathcal D}'
= {\mathcal D} - D$.
\begin{lem}\label{lem3.1} Assume $\nabla_{/K}|C-D$ lifts to an
absolute integrable connection $\nabla' : L_{C-D} \to
L_{C-D}\otimes\Omega^1_{C-D/k}$. Then $\nabla'$ extends to an
absolute integrable connection
\eq{\nabla : L \to L\otimes\Omega^1_{C/k}\langle D \rangle({\mathcal D}').
}{3.1}
The notation $\langle D \rangle$ refers to log poles at $D$ as in \eqref{1.4}.
\end{lem}
\begin{proof}Let $e$ be a basis for $L$ at $c$ a point with
multiplicity $m\ge 1$ in ${\mathcal D}$, and let $x$ be a
local parameter at $c$ on $C$. Write
\eq{\nabla'(e) = A(x)dx+\sum_iB_i(x)d\tau_i;\quad d\tau_i \text{ basis
in } \Omega^1_K,\
x^mA(x)\in {\mathcal O}_{C,c}.
}{3.2}
We must show $x^{m-1}B_i(x)$ is regular at $c$. But integrability of
$\nabla'$ implies that $\partial A/\partial \tau_i = \partial B_i/\partial x$, from which the
assertion is clear.
\end{proof}
We know from proposition \ref{prop2.16}
that the restriction to $C-D$ of an integrable
absolute connection of the form
\eqref{3.1} pulls back from a unique invariant integrable absolute
connection ${\mathcal L} \to {\mathcal L}\otimes\Omega^1_G$ on $G=J_{\mathcal D}$. More precisely, we fix a
basepoint $c_0 \in (C-D)(K)$ and normalize our connection \eqref{3.1}
to be trivial at the basepoint by tensoring with a pullback from
${\rm Spec \,}(K)$.
We consider now the basic geometric picture of Deligne \cite{De}
\eq{\pi:\text{Sym}^N(C-D) \to G_N;\quad N=2g-2+\sum_im_i
}{3.3}
where $G_N$ is the $J_{\mathcal D}$-torseur of degree $N$ line bundles
trivialized along ${\mathcal D}$ and $\pi(\sum z_i)={\mathcal O}(\sum z_i)$ with
trivialization given by restricting to ${\mathcal D}$ the canonical (upto
scalar in $K$) map ${\mathcal O}_C \to {\mathcal O}_C(\sum z_i)$. Note that $N=\deg(\Omega^1_{C/K}({\mathcal D}))$
and $\dim G = g-1+\sum m_i = N-g+1$. (Recall we assume ${\mathcal D}\neq
\emptyset$.) We identify $G_N \cong G$ by sending the point
$[{\mathcal O}(Nc_0)]\mapsto 0$, and we write ${\mathcal L}$ for the resulting line
bundle with connection on $G_N$. The basic remark of Deligne is
\begin{prop}\label{prop3.2}Assume
$$H^0_{DR}(C-D,(L,\nabla))=H^2_{DR}(C-D,(L,\nabla)) =
0.$$
Then
\ml{\det(H^*_{DR}(C-D,(L,\nabla))=\det(H^1_{DR}((C-D,(L,\nabla)))) \\
\cong H^N(\text{Sym}^N(C-D),(\pi^*({\mathcal L}, \nabla)))
}{3.4}
as a line with connection on $K$.
\end{prop}
\begin{proof} Our hypotheses imply $H^1_{DR}(C-D,(L,\nabla))$ has
dimension $N$. Consider the diagram
\eq{\begin{array}{ccc} (C-D)^N & \stackrel{p}{\to} &
\makebox[2.4cm][r]{$\text{Sym}^N(C-D)$} \\
& \searrow q & \makebox[2.4cm][l]{$\downarrow \pi$} \\
&& \makebox[2.4cm][l]{$G_N$}
\end{array}
}{3.5}
We have $q^*({\mathcal L}) \cong L\boxtimes\cdots\boxtimes L$ (exterior tensor
product on $(C-D)^N$. The K\"unneth formula gives
\eq{H^N((C-D)^N,(L\boxtimes\cdots\boxtimes L,\nabla)) \cong
H^1_{DR}(C-D,(L,\nabla))^{\otimes N}.
}{3.6}
There is an action of the symmetric group ${\mathcal S}_N$ on the pair
$$((C-D)^N, L\boxtimes\cdots\boxtimes L).$$ The resulting action on
$(H^1_{DR})^{\otimes N}$ is alternating because of the odd degree
cohomology, so the invariants are precisely $\det H^1_{DR}$. There is
an evident map
\ml{p^*: H^N_{DR}(\text{Sym}^N(C-D),\pi^*{\mathcal L}) \to
H^N_{DR}((C-D)^N,L\boxtimes\cdots\boxtimes L)^{{\mathcal S}_N} \\
= \det H^1_{DR}(C-D,L)
}{3.7}
To show this map is an isomorphism, it suffices to remark that one has
a trace map
\eq{p_* : H^N_{DR}((C-D)^N,L\boxtimes\cdots\boxtimes L) \to
H^N_{DR}(\text{Sym}^N(C-D),\pi^*{\mathcal L})
}{3.8}
Because $L\boxtimes\cdots\boxtimes L = p^*\pi^*{\mathcal L}$, the existence of
such a trace follows from the projection formula and the trace in de
Rham cohomology with constant coefficients.
\end{proof}
Now one uses the geometry of the map $\pi$ and \eqref{3.4} to compute
the determinant.
\begin{lem}\label{lem3.3} Let $X$ be a smooth variety over a field of
characteristic $0$. Let $A\subset X$ be a smooth subvariety of
codimension $p$. Let $(E,\nabla)$ be an integrable connection on
$X$. Then
\eq{\H^n_A(X, E\otimes \Omega^*_X) \cong H^{n-2p}(A, E\otimes
\Omega^*_A).
}{3.9}
\end{lem}
\begin{proof} Write $\underline{\underline{H}}_A^r(F)$ for the Zariski sheaf associated to the
presheaf $U \mapsto H^r_A(U,F)$ for any Zariski sheaf $F$ on $X$. For
$F$ locally free, $\underline{\underline{H}}_A^r(F)= (0)$ for $r\neq p$ by purity. Duality
theory gives (here $X\supset A_\alpha \supset A$
runs through
nilpotent thickenings)
\ml{E\otimes \Omega^m_A \to {\rm Ext}^p({\mathcal O}_A, E\otimes \Omega^{m+p}_X)
\to \varinjlim_\alpha {\rm Ext}^p({\mathcal O}_{A_\alpha},E\otimes\Omega^{m+p}_X) \\
\cong \underline{\underline{H}}_A^p(E\otimes\Omega^{m+p}_X).
}{3.10}
We want to show that this map is an isomorphism, compatible
with the connection, thus yielding a quasiisomorphism
of complexes
\eq{E\otimes\Omega^*_A \to \underline{\underline{H}}_A^p(E\otimes\Omega^*_X).
}{3.11}
The problem is local, so we can assume
$$A\subset A_1 \subset \ldots \subset A_p = X
$$
with $A_i$ smooth of codimension $p-i$ in $X$. Now
$\underline{\underline{H}}_A^p(E\otimes\Omega^*_X)$ represents
$R\underline{\underline{\Gamma}}_A(E\otimes\Omega^*_X)[p]$ in the
derived category of Zariski sheaves on $A$, and in the derived
category we may write
\eq{R\underline{\underline{\Gamma}}_A(E\otimes\Omega^*_X)[p] =
R\underline{\underline{\Gamma}}_A[1]\circ
R\underline{\underline{\Gamma}}_{A_1} [1]\circ\ldots\circ
R\underline{\underline{\Gamma}}_{A_{p-1}}(E\otimes\Omega^*_X)[1]
}{3.12}
In this way, we reduce to verifying \eqref{3.9} in the case $p=1$. So,
suppose $A:t=0$ in $X={\rm Spec \,}(R)$. We have
\eq{ \underline{\underline{H}}_A^1(E\otimes\Omega^*_X)\cong
E_{R[t^{-1}]}\otimes
\Omega^*_{R[t^{-1}]}/E_R\otimes\Omega^*_R
}{3.13}
as $ H^1_{DR}(X, E) \subset H^1_{DR}(X-A, E).$
By \cite{DeI},
since $E$ has no singularity along $t=0$, one has
\eq{(E_{R}\otimes \Omega^*_{R}(\log(t=0)), \nabla)
\stackrel{{\rm q.iso.}}{\to}
(E_{R[t^{-1}]}\otimes \Omega^*_{R[t^{-1}]}, \nabla|_{{\rm Spec}R[t^{-1}]})
}{3.14}
Thus
${\rm res \ }: \underline{\underline{H}}_A^1(E\otimes\Omega^*_X, \nabla)) \to
E_{R/tR} \otimes (\Omega^*_{R/tR}[-1], \nabla|_{{\rm Spec}(R/tR)})
$
is an isomorphism.
\end{proof}
\begin{lem}\label{lem3.4}
Let $p:G_N \to J_N$ be the projection to the corresponding
torseur over the absolute jacobian. Write $[\Omega^1_{C/K}({\mathcal D})]\in J_N$ for the
point corresponding to the canonical bundle twisted by ${\mathcal O}({\mathcal D})$. Let
$a\in G_N$. We have
\eq{\pi^{-1}(a) = \begin{cases} {\mathbb A}^{g-1} & p(a) \neq [\Omega^1_{C/K}({\mathcal D})] \\
{\mathbb A}^g & p(a) = [\Omega^1_{C/K}({\mathcal D})];\ \partial(a) = 0 \\
\emptyset & p(a) = [\Omega^1_{C/K}({\mathcal D})];\ \partial(a) \neq 0
\end{cases}
}{3.15}
where by definition ${\mathbb A}^{g-1}= \emptyset$ is $g=0$.
Note that if $p(a)=[\Omega^1_{C/K}({\mathcal D})]$, then $a$ corresponds to a
trivialization ${\mathcal O}_{\mathcal D} \cong \Omega^1_{C/K}({\mathcal D})|_{\mathcal D}$ defined upto
scalar. $\partial(a)$ in the above refers to the evident boundary of this
trivialization in $H^1(\Omega^1_{C/K})=K$ (again upto scale).
\end{lem}
\begin{proof}Let $p(a)$ correspond to a line bundle $M$ of degree $N$,
we consider the exact sequence
\eq{0 \to M(-{\mathcal D}) \to M \to M|_{\mathcal D} \to 0
}{3.16}
Suppose first $M\neq \Omega^1_{C/K}({\mathcal D})$. Then $H^1(M(-{\mathcal D}))=(0)$, so any
trivialization in $H^0(M|_{\mathcal D})$ lifts to $H^0(M)$, and the space of
such liftings is a torseur under $H^0(M(-{\mathcal D}))$, a vector space of
dimension $g-1$. (Note this is an affine torseur, not a projective
torseur.) If $M = \Omega^1_{C/K}({\mathcal D})$, $H^0(M(-{\mathcal D}))$ has dimension $g$, and
the image $H^0(M) \to H^0(M|_{\mathcal D})$ has codimension $1$.
\end{proof}
\begin{remark}\label{rmk3.5} If we choose local parameters $t_i$ at
$c_i \in {\mathcal D}$, then
$$H^0(\Omega^1_{C/K}({\mathcal D})|_{\mathcal D})$$
can be identified with the
space of polar parts of $1$-forms with poles along ${\mathcal D}$, and the map
$\partial$ is given by the residue
\eq{\partial(\sum_i\sum_{j=0}^{m_i-1} u_{ij}dt_i/t_i^{m_i-j}) = \sum_i
u_{i,m_i-1}
}{3.17}
Note the (open) condition for an element in $H^0(\Omega^1_{C/K}({\mathcal D})|_{\mathcal D})$ to
be a trivialization is simply
\eq{\prod_i u_{i0} \neq 0.
}{3.18}
Because $B\subset G_N$, we must factor out by the action of ${\mathbb G}_m$,
which we can normalize away by setting $u_{10} = 1$.
Thus we have
\ml{B:= \pi(\text{Sym}^N(C-D))\cap p^{-1}[\Omega^1_{C/K}({\mathcal D})] = \\
\Big\{\sum_i\sum_{j=0}^{m_i-1} u_{ij}dt_i/t_i^{m_i-j} \ \Big| \
\sum_i u_{im_i-1}= 0;\
\prod_i u_{i0} \neq 0;\ u_{10}=1 \Big\}.
}{3.19}
\end{remark}
Define $A:= \pi^{-1}(B)\subset \text{Sym}^N(C-D)=:X$. Using the localization
sequence and lemma \ref{lem3.4}, we get an exact sequence
\eq{\minCDarrowwidth.5cm \begin{CD}
H^{N-2g}_{DR}(A/K,\pi^*{\mathcal L}) @>>> H^N_{DR}(X/K,\pi^*{\mathcal L}) @>>>
H^N_{DR}(X - A/K,\pi^*{\mathcal L}) \\
@AA\cong A @. @AA\cong A \\
H^{N-2g}_{DR}(B/K,{\mathcal L}|_B) @.@. \makebox[3cm][r]{$H^N_{DR}(G_N -
p^{-1}[\Omega^1_{C/K}({\mathcal D})]/K,{\mathcal L})$}
\end{CD}
}{3.20}
where $p:G_N \to J_N$ with $J= J(C)$ the absolute Jacobian. The fact
that the vertical arrows in this diagram are isomorphisms follows
because the maps are maps of affine space bundles and the line bundles
with connection are pulled back from the base.
To simplify the presentation, we will assume that $m_1\ge 2$. Another
proof of our formula for the de Rham determinant in the case ${\mathcal D} = D
= \sum c_i$, (i.e. for regular singular points) will be given in
theorem \ref{thm4.5}.
\begin{lem}\label{lem3.6} Assume the line bundle $L$ on $C$ has degree
$0$, and that ${\mathcal D}$ is minimal, i.e. $\nabla: L|_{\mathcal D} \cong
L\otimes\Omega^1_{C/K}({\mathcal D})|_{\mathcal D}$. We continue to assume also that ${\mathcal D}=\sum
m_ic_i$ with $m_1\ge 2$. Then
\eq{H^*_{DR}(\text{Sym}^N(C-D)-A/K,\pi^*{\mathcal L}) = (0).
}{3.21}
\end{lem}
\begin{proof}The isomorphism on the right in \eqref{3.20} implies we
must show $H^*_{DR}(G_N - p^{-1}[\Omega^1_{C/K}({\mathcal D})]/K,{\mathcal L}) = (0)$. The
assumption $m_1\ge 2$ means we have a ${\mathbb G}_a$ action by translation on
$ G_N - p^{-1}[\Omega^1_{C/K}({\mathcal D})]/K,{\mathcal L} $, and minimality of ${\mathcal D}$ implies
that the connection is nontrivial on the fibres. The fibration is
Zariski-locally trivial, so the Leray spectral sequence for de Rham
cohomology reduces us to showing $H^*_{DR}({\mathbb G}_{a,S}/S,({\mathcal O},\Xi))=(0)$
where $\Xi$ is an everywhere non-zero, translation invariant, relative
$1$-form on
${\mathcal O}_{{\mathbb G}_{a,S}}$. In other words, for $s\in {\mathcal O}_S^\times$ we must show
\eq{{\mathcal O}_S[t] \stackrel{d+sdt}{\longrightarrow} {\mathcal O}_S[t]dt
}{3.22}
has trivial cohomology. This is straightforward.
\end{proof}
\begin{lem}\label{lem3.7} Assume $H^0_{DR}(C-D,L)=(0)$. Then
\eq{H^{m-2}_{DR}(B/K,{\mathcal L}|_B) \cong
H^N_{DR}(\text{Sym}^N(C-D)/K,\pi^*{\mathcal L})\cong K
}{3.23}
as a line with a connection over $K$.
\end{lem}
\begin{proof}Note $m-2=N-2g$. Extending the top sequence in
\eqref{3.20} one step to
the left and using the previous lemma gives the left isomorphism. We
have already seen the isomorphism on the right.
\end{proof}
Our task now is to calculate $H^{m-2}_{DR}(B/K,{\mathcal L}|_B)$ with its
connection. We assume that ${\mathcal L} \in {\rm Pic}^0(J)$ as in lemma
\ref{lem3.6}. Then ${\mathcal L}$ carries a relative invariant connection
$d_{/K}$ on $J$, and
$\nabla_{/K}= d_{/K}+ \Xi$ for some invariant form
$\Xi \in H^0(G, \Omega^1_{G/K})^{{\rm inv}}$.
Changing the choice of $d_{/K}$ changes $\Xi$
to $\Xi + p^*(\alpha)$, where $\alpha \in H^0(J, \Omega^1_{J/K})^{{\rm
inv}}$ and $p: G_N \to J_N$ is the torseur under the affine
group ${\mathcal G}:= \ker(G \to J)$.
In particular, $\Xi|p^{-1}[\Omega^1_{C/K}({\mathcal D})]$ does not
depend on the choice of $d_{/K}$.
As $p^{-1}(\Omega^1_{G/K}({\mathcal D}))$ is isomorphic to ${\mathcal G}$, and
we see that
\eq{({\mathcal L},\nabla_{/K})|_B = ({\mathcal O}_B, d+ \Xi|_B).
}{3.24}
We say that $\Xi|_B$ vanishes at a point $b\in B$ if $\Xi(b)=0$ in
$\mathfrak
m_b/\mathfrak m_b^2$.
\begin{lem}\label{lem3.8} Let $b\in B\subset G_N$ correspond to the
trivialization on $\Omega^1_{C/K}({\mathcal D})|_{\mathcal D}$ given by $\nabla|_{\mathcal D} : L|_{\mathcal D} \to
L\otimes \Omega^1_{C/K}({\mathcal D})|_{\mathcal D}$. Then $\Xi|_B$ vanishes at $b$. $\Xi|_B$ does
not vanish at any other point of $B$.
\end{lem}
\begin{proof}
Let us write $\eta:=i^*\Xi|C-D$, where $i: C-D \to G$ is the cycle
map. Then by definition, the trivialization
of $\Omega^1_{C/K}({\mathcal D})|_{\mathcal D}$
associated to
$i^*({\mathcal L},\nabla)$ depends only on $\eta|{\mathcal D}$, or equivalently
only on $\Xi|p^{-1}[\Omega^1_{C/K}({\mathcal D})]$.
We have
\eq{\pi^*\Xi = \sum_{i=1}^N \eta_i
}{3.25}
where $\eta_i$ is the pullback of $\eta$ via the $i$-th projection
$(C-D)^N \to C-D$.
Suppose for a moment that the divisor of $\eta$ (viewed as a section
of $\Omega^1_{C/K}({\mathcal D})$) is reduced, $(\eta) = \sum e_i;\ e_i \in
(C-D)(\bar{K})$. Let $e := (e_1,\cdots,e_N)\in \text{Sym}^N(C-D)$ be the
point corresponding to $\eta$. We have $e\in A=\pi^{-1}(B) \subset
\text{Sym}^N(C-D) $ and $b=\pi(e)$. Since $A \to B$ is a projective
bundle, there is a
surjection on tangent spaces
\eq{\begin{array}{ccc}T_A(e) & \hookrightarrow & T_{\text{Sym}^N(C-D)}(e) \\
\makebox[0cm][r]{surjective}\downarrow && \downarrow \pi_* \\
T_B(b) & \hookrightarrow & T_{G_N}(b) \end{array}
}{3.26}
Since $T_{\text{Sym}^N(C-D)}(e)$ is spanned by expressions $\sum
\tau_i|_{e_i}$, to show $\Xi|_B$ vanishes at $b$, it suffices to show
\eq{<\Xi,\pi_*(\sum \tau_i|_{e_i})> = 0.
}{3.27}
This expression equals
\eq{<\pi^*(\Xi),\sum \tau_i|_{e_i}> = \sum <\eta,\tau_i|_{e_i}>.
}{3.28}
Each term on the right vanishes because $\eta(e_i)=0$ in $\mathfrak
m_{e_i}/\mathfrak m_{e_i}^2$. The general case ($(\eta)$ not necessarily
reduced) follows from this by a specialization argument.
We postpone until lemma \ref{lem3.10}
the proof that $\Xi|_B$ doesn't vanish at any
other point of $B$.
\end{proof}
By assumption we start with an absolute, invariant, integrable
connection on ${\mathcal L}$ of degree 0 on $J$. Restricting to $B$, we get an
absolute closed
invariant $1$-form $\Psi$, whose corresponding relative form is
$\Xi$. Recall \eqref{3.19} we have coordinates $u_{ij}$ on $B$ with
$u_{10}=1$, $\prod u_{i0}\neq 0$, and $u_{1,m_1-1}=-\sum_{i\ge
2}u_{i,m_i-1}$.
\begin{lem}\label{lem3.9}
Under the assumption that
$$\{i^*\nabla_{/K}: i^*{\mathcal L} \to i^*{\mathcal L} \otimes
\Omega^1_{C/K}({\mathcal D})\} \to \{j_*i^*{\mathcal L}|(C-D)
\to j_*i^*{\mathcal L} \otimes \Omega^1_{(C-D)/K}\}$$
is a quasiisomorphism, and $i^*\nabla_{/K}$ has poles
along all points of $D$,
we can arrange that a $K$-basis for
$$H^{m-2}_{DR}(B/K,{\mathcal L}|_B) = H^{m-2}_{DR}(B/K,({\mathcal O}_B,d+\Xi))
$$
is given by the closed form
\eq{\theta:= \prod_{\substack{(i,j) \\ m_i\ge 2}}
du_{ij}\wedge\prod_{\substack{i \\ m_i=1}}
\frac{du_{i0}}{u_{i0}}
}{3.29}
\end{lem}
\begin{proof} Recall \eqref{2.29} the relative invariant forms on
${\mathcal G}:=\ker(G \to J)$ are the $\tau_{ij}$ defined by
the expression
\eq{\sum_{j=0}^{m_i-1} \tau_{ij}T_i^j = \Big(\sum_j
u_{ij}T_i^j\Big)^{-1} \sum_j du_{ij}T_i^j.
}{3.30}
Write
\eq{\Xi = \sum_{i,j}\lambda_{ij}\tau_{ij};\quad \lambda_{i,m_i-1}\neq 0
}{3.31}
where the nonvanishing condition comes from the requirement that the
form restricted to $C-D$ gives a trivialization along ${\mathcal D}$
(see \eqref{2.30}). If we
write ($\mod T_i^{m_i}$)
\eq{\Big(u_{i0}+u_{i1}T_i+\ldots+u_{i,m_i-1}T_i^{m_i-1}\Big)^{-1} =
\nu_{i0}+ \nu_{i1}T_i+\ldots+\nu_{i,m_i-1}T_i^{m_i-1}
}{3.32}
we get the table
\begin{align}\label{3.33}\tau_{i0} & = \nu_{i0}du_{i0} \\
\notag\tau_{i1} & = \nu_{i0}du_{i1}+\nu_{i1}du_{i0} \\
\notag\vdots \\
\notag\tau_{i,m_i-1} & = \nu_{i0}du_{i,m_i-1}+\ldots+\nu_{i,m_i-1}du_{i0}
\end{align}
Note if we give $u_{ij},\ du_{ij},\ \nu_{ij}$ all weight $j$, then
$\tau_{ij}$ will be homogeneous of weight $j$.
Comparing \eqref{3.33} and \eqref{3.31}, it follows that if we expand
$\Xi|_B$ in terms of the $du_{ij}$, omitting $du_{10}$ and
$du_{1,m_1-1}$, we find for suitable $\alpha_{ij} \neq 0$
\ml{\Xi|_B = \sum_i g_{ij}du_{ij} = \sum_{i,j}
\Big[(\alpha_{i,m_i-1}u_{i0}^{-1}-\alpha_{1,m_1-1})du_{i,m_i-1}+ \\
u_{i0}^{-1}\sum_{p=0}^{m_i-2}
\Big(\alpha_{ip} \frac{u_{i,m_i-1-p}}{u_{i0}}+
\sum \text{ terms at least
quadratic in $\frac{u_{ik}}{u_{i0}}$} \Big)du_{ip}\Big]
}{3.34}
Looking at the weights, we see that for $p\le
m_i-2$
\ml{g_{ip} = \text{ nonzero multiple of
}\frac{u_{i,m_i-p-1}}{u_{i0}}+ \\
\text{ terms only involving $u_{ij},\
j<m_i-p-1$}
}{3.35}
while
\eq{g_{i,m_i-1} = \alpha_{i,m_i-1}u_{i0}^{-1}-\alpha_{1,m_1-1}
}{3.36}
with neither $\alpha$ coefficient $0$.
Now generators of $H^{m-2}_{DR}(B,({\mathcal O},d+
\Xi))$ are of the form
$M\theta$ where $M$ is a monomial in the $u_{ij}, u_{i0}^{-1}$ and
$\theta$ is as in \eqref{3.29}. Relations are
\eq{\Big(\frac{\partial}{\partial u_{ij}}+g_{ij}\Big)(M)\theta = 0
}{3.37}
Because of \eqref{3.35} one can use these relations to eliminate
$u_{ij},\ j>0$ from $M$ by downward induction on $j$, starting from
$u_{i,m_i-1}$. We are left with the case $M=u_{2,0}^{n_2}\cdots
u_{r,0}^{n_r}$ with $n_i\in {\mathbb Z}$. In this case we can apply
\eqref{3.36}. If $m_i\ge 2$, we get the relation
\eq{u_{2,0}^{n_2}\cdots u_{r,0}^{n_r}\theta \equiv
\frac{\alpha_{i,m_i-1}}{\alpha_{1,m_1-1}}u_{2,0}^{n_2}\cdots
u_{i,0}^{n_i-1}\cdots u_{r,0}^{n_r} \theta
}{3.38}
Using this, we can get $n_i=0$. If $m_i=1$ and $i\ge 2$ the relation
becomes
\eq{u_{2,0}^{n_2}\cdots u_{r,0}^{n_r}\theta \equiv
\frac{n_i+\alpha_{i,0}}{\alpha_{1,m_1-1}}u_{2,0}^{n_2}\cdots u_{i,0}^{n_i-1}\cdots
u_{r,0}^{n_r} \theta
}{3.39}
If $\alpha_{i,0}$ is not a positive integer, we can arrange $n_i=0$.
On the other hand, we claim that if $m_i=1$, then
$\alpha_{i0}$ is the residue of $i^*\nabla$ along $c_i$.
Indeed \eqref{2.29} shows that in this case,
$\tau_{i0}= d\log u_{i0}$, and $u_{i0}$ is then just the local
parameter in the point $c_i$ (see \eqref{2.28}).
Thus the quasiisomorphism
\begin{multline}\label{3.40}
\{i^*\nabla_{/K}: i^*{\mathcal L} \to i^*{\mathcal L} \otimes
\Omega^1_{C/K}({\mathcal D})\} \to \{j_*i^*{\mathcal L}|(C-D) \\
\to j_*i^*{\mathcal L} \otimes \Omega^1_{(C-D)/K}\} \end{multline}
forces $a_i$ not to lie in ${\mathbb N} -\{0\}$.
\end{proof}
The following was left open in the proof of lemma \ref{lem3.8}:
\begin{lem}\label{lem3.10}Let $\Xi=\sum_{i,j}\lambda_{ij}\tau_{ij}$ be
as in lemma \ref{lem3.8}. Then $\Xi$ vanishes at a unique point $b\in
B$.
\end{lem}
\begin{proof}We have seen in the proof of lemma \ref{lem3.8} that
$\Xi$ vanishes at a point in $B$. We must show it vanishes at at most
one point. Let $b=(\ldots,b_{ij},\ldots)\in B$ be a point. Write
$b=(\ldots,y_{ij},\ldots)$ with respect to the coordinates $\nu_{ij}$
\eqref{3.32}. Staring at \eqref{3.33}, the conditions that $\Xi|_b=0$
are seen to be (recall $\sum_i du_{i,m_i-1}=0=du_{10}$) for $i\ge 2$
\begin{gather}
\lambda_{i,m_i-1}y_{i1}+\lambda_{i,m_i-2}y_{10} =0 \label{3.41}\\
\lambda_{i,m_i-1}y_{i2}+\lambda_{i,m_i-2}y_{i1}+\lambda_{i,m_i-3}y_{i0}
=0 \notag\\
\vdots \ \ \vdots \notag\\
\lambda_{i,m_i-1}y_{i,m_i-1}+\lambda_{i,m_i-2}y_{i,m_i-2}+\ldots +
\lambda_{i,0}y_{i0} = 0 \notag
\end{gather}
For $i=1$ one gets the same list but with the last line (coefficient
of $du_{i0}$) omitted. Finally, using $\nu_{10}=1$ and
$du_{1,m_1-1}=-\sum_{i\ge 2} du_{i,m_i-1}$ one gets
\eq{\lambda_{1,m_1-1}=y_{i0}\lambda_{i,m_i-1};\quad 2\le i\le r
}{3.42}
Since $\lambda_{i,m_i-1}\neq 0$, equations \eqref{3.41} and
\eqref{3.42} admit a unique solution for the $y_{ij}$. Since we know
$\Xi$ vanishes at at least one point of $B$, this point must lie in
$B$.
\end{proof}
Finally, we must calculate the Gau\ss-Manin connection on
$$H^{m-2}_{DR}(B,({\mathcal O}_B,d +\Xi))$$
Define $\Psi$ to be an absolute
invariant form lifting $\Xi$. By assumption our connection on $C-D$
comes from an absolute integrable connection which, by proposition
\ref{prop2.16}, comes from an absolute integrable connection on
$G$. Restricting this connection to $B$ gives our $\Psi$.
\begin{lem}\label{lem3.11}With notation as above, there exists $F\in
{\mathcal O}_B,\ \eta\in {\Omega^1_K}$, and $a_i\in k,\ i\ge 2$, such that
\eq{\Psi = \sum_{i=2}^r a_i\frac{du_{i0}}{u_{i0}}+dF+\eta.
}{3.43}
If moreover $\nabla$ is integrable, then $\eta$ is closed.
\end{lem}
\begin{proof}Since $\Xi$ is (relatively) closed on $B$, one can write
\eq{\Xi = \sum_{i\ge 2} a_i\frac{du_{i0}}{u_{i0}}+d_{/K}F;\quad a_i \in K.
}{3.44}
Lifting to an absolute form forces
\eq{\Psi = \sum_{i\ge 2} a_i\frac{du_{i0}}{u_{i0}}+dF + \sum_j
f_j\eta_j;\quad f_j\in {\mathcal O}_B,\ \eta_j\in \Omega^1_K.
}{3.45}
Here the $\eta_j$ are linearly independent in $\Omega^1_K$.
Using $d\Psi=0$ modulo $\Omega^2_K \otimes {\mathcal O}_B$
and taking residues along $u_{i0}=0$ yields $a_i\in
k\subset K$. Then computing $d\Psi \mod {\mathcal O}_B\otimes\Omega_K^2$ yields
\eq{0=\sum_j d_{/K}f_j\otimes\eta_j\in \Omega^1_{B/K}\otimes\Omega^1_K.
}{3.46}
It follows that $f_j\in K$, so $\eta:=\sum f_j\eta_j\in \Omega^1_K$. Taking
$d$ again shows $\eta$ is closed if $\nabla$ is integrable.
\end{proof}
We now compute the Gau\ss-Manin connection. We have the diagram of
global sections
\eq{\minCDarrowwidth.5cm \begin{CD}@. \Omega^{m-2}_{B/k}
@>>\text{onto} >
\Omega^{m-2}_{B/K} \\
@. @VV d+\Psi V \\
\Omega^{m-2}_{B/K}\wedge \Omega^1_K @>>\cong >
\frac{\Omega^{m-1}_{B/k}}{\Omega^{m-3}_{B/k}\cdot \Omega^2_K} @.
\end{CD}
}{3.47}
The connection is determined by its value on $\theta$ \eqref{3.29}. To
calculate, one lifts $\theta$ to $\tilde\theta\in \Omega^{m-2}_{B/k}$
and then applies $d+\Psi$. But for $\tilde\theta$ one can choose the
form with the same expression \eqref{3.29}. This form is closed, so
\eq{\nabla_{GM}(\theta) = \Psi\wedge\theta = (d_K(F)+\eta)\wedge\theta
}{3.48}
Here we write $F=\sum_I a_Iu^I,\ a_i\in K$ and $d_K(F) := \sum
da_Iu^I$.
Let $b\in B$ be the point corresponding to the trivialization of
$\Omega^1_{C/K}({\mathcal D})$ given by the polar part of
the original relative connection. It really lies in $B$ since we
have assumed that ${\rm deg} {\mathcal L}=0$.
\begin{lem}\label{lem3.12} With notation as above, the
Gau{\ss}-Manin connection on the rank $1$ $K$-vector space
$$H^{m-2}_{DR}(B,({\mathcal O},d+\Xi))$$
described by \eqref{3.48} is isomorphic
to the connection on $K$ given by
$$1\mapsto \Psi|_b + \frac{1}{2}d\log(\kappa)
$$
for a suitable $\kappa\in K^\times$.
\end{lem}
\begin{proof}We have seen (lemma \ref{lem3.8}) that this point $b$ is
determined by the condition that $\Xi(b)=0\in \mathfrak
m_{B,b}/\mathfrak m_{B,b}^2$. Changing $\Psi$ by a closed form pulled back
from $K$ changes the Gau{\ss}-Manin connection and the connection at $b$
in the
same way, so we can assume $\eta=0$, i.e. $\Psi = \sum
a_i\frac{du_{i0}}{u_{i0}} +dF$. Write
\eq{g_{ij} = \begin{cases}\frac{\partial
F}{\partial u_{ij}} & j>0 \\ \frac{\partial
F}{\partial u_{i0}}+a_i/u_{i0} & j=0. \end{cases}
}{3.49}
Write $F = \sum_I a_I u^I$. Then
\eq{\Psi = \sum g_{ij}du_{ij}+\sum_I
u^Ida_I;\quad \Psi\wedge\theta = \sum_I u^Ida_I\wedge\theta
}{3.50}
We have
\eq{g_{ij}(b) = 0,\ j<m_i-1;\quad g_{i,m_i-1}(b) = g_{k,m_k-1}(b);\
\text{all }i,k.
}{3.51}
Since $\sum_i du_{i,m_i-1}|_B=0$, we see from \eqref{3.51} that
\eq{\Psi|_{\{b\}} = \sum_Ib^Ida_I
}{3.52}
Thus, it will suffice to relate $u^I\theta$ and $b^I\theta$ in
$H^{m-2}_{DR}$. Note that each monomial $u^I$ involves $u_{ij}$ for
only one value of $i$, and the weight of $u^I$ is $\le m_i-1$ (see the
discussion after \eqref{3.33}).
Suppose first the weight of $u^I$ is strictly less than $m_i-1$. Let
$j$ be maximal such that $u_{ij}$ appears in $u^I$. From \eqref{3.35}
it follows that
\eq{g_{i,m_i-1-j} = \alpha_{i,m_i-1-j}\frac{u_{ij}}{u_{i0}^2} +
\text{terms involving only } u_{ik};\ k<j.
}{3.53}
Here $\alpha_{i,m_i-1-j}\neq 0$. Define $u^L =
u^Iu_{i0}^2u_{ij}^{-1}$. Note the weight of $u^L$ is $<m_i-1-j$, so in
$H^{m-2}_{DR}$ we have (compare \eqref{3.37})
\ml{u^I\theta = (u^I - \alpha_{i,m_i-1-j}^{-1}(\frac{\partial}{\partial
u_{i,m_i-1-j}}+g_{i,m_i-1-j})u^L)\theta = \\
(u^I - \alpha_{i,m_i-1-j}^{-1}g_{i,m_i-1-j}u^L)\theta = \sigma_I\theta
}{3.54}
where $\sigma_I$ is a sum of terms
of weights $< |I|$ and terms of weight $|I|$ only involving
$u_{i0},\dotsc,u_{i,j-1}$. Note that $b^I= \sigma_I(b)$ because
$g_{i,m_i-1-j}(b) = 0$. In this way we reduce to the case
$u^I=u_{i0}^p$. Our assumption on the weight implies $m_i\ge 2$, so
\eq{(\frac{\partial}{\partial u_{i,m_i-1}}+g_{i,m_i-1})u_{i0}^p =
g_{i,m_i-1}u_{i0}^p.
}{3.55}
Together with \eqref{3.35} and $g_{ij}(b)=0$, this enables us to
reduce to $p=0$.
Suppose now the weight of $I$ is
$m_i-1$. If $m_i\ge 2$ we can use the above argument, except in the
case $u^I = u_{ij}u_{i,m_i-1-j}u_{i0}^{-2}$. Here there are two subcases.
If $j\neq m_i-1-j$, the
$\alpha_{i,m_i-1-j}$ in
\eqref{3.54} is $a_I$ in the expansion $F=\sum_I a_I u^I$, so
\eq{u^Ida_I\wedge\theta = (b^Ida_I+\frac{da_I}{a_I})\wedge\theta.
}{3.56}
This completes the proof in this case because the connections
$b^Ida_i$ and $b^Ida_I+d\log(a_I)$ are isomorphic. If, on the other hand,
$m_i$ is odd and $j=\frac{m_i-1}{2}$, the monomial $u^I =
u_{ij}^2u_{i0}^{-2}$ and $dF$ contains the term $2a_Iu_{ij}du_{ij}$.
Thus, from \eqref{3.53} we conclude $\alpha_{i,\frac{m_i-1}{2}}= 2a_I$.
The lefthand identity in \eqref{3.54} yields in this case
\eq{u^Ida_I\wedge\theta = (b^Ida_I + \frac{1}{2}d\log
a_I)\wedge\theta.
}{sp}
In the statement of the lemma, we take $\kappa$ to be the product of the
corresponding $a_I$.
Suppose finally $m_i=1$. In this case $\frac{\partial F}{\partial
u_{i0}}=0$, so the corresponding $a_I=0$ and by \eqref{3.52} this term
contributes nothing to $\Psi|_{\{b\}}$. Similarly, by \eqref{3.50}
there is no contribution to $\Psi\wedge\theta$.
\end{proof}
We give two interpretations of the $2$-torsion term
$\frac{1}{2}d\log(\kappa)$ occurring in the previous lemma.
\begin{defn}\label{defins1} Let $\sigma$ be a closed $1$-form relative to
$K$ on
$${\rm Spec \,}(K[[t_1,\dotsc,t_N]]).
$$
Assume $\sigma(0)=0\in \frak m/\frak m^2$.
Write $\sigma = dh$ with $h(0)=0$, so $h=h_2+h_3+\ldots$ with $h_i$
homogeneous of degree $i$. If $h_2$ is nondegenerate, we may define
$\text{disc($\sigma)$}=
\text{discriminant}(h_2)\in K^{\times}/K^{\times 2}$. This is
well-defined independent of the choice of parameters.
\end{defn}
\begin{thm} The Gau{\ss}-Manin connection on
$H^{m-2}_{DR/K}(B,({\mathcal O},d+\Xi))$ is isomorphic to
$$(d+\Psi)|_{\{b\}} +
\frac{1}{2}d\log({\rm disc}(\Xi|_{\widehat{{\mathcal O}}_{B,b}})).
$$
(In particular, the quadratic term in $h=\int
\Xi|_{\widehat{{\mathcal O}}_{B,b}}$is non-degenerate.)
\end{thm}
\begin{proof}First we collect some facts about $\Xi =
\sum_{i,j}\lambda_{ij}\tau_{ij}= \sum g_{ij}du_{ij}$. We have $u_{1,0}=1$
and $\nu_{i,0} = u_{i,0}^{-1}$. It follows from \eqref{3.33} of the paper
that
$u_{1,m_1-1}$ does not appear in the expression for $\Xi$ and
$du_{1,m_1-1}$ only appears with constant coefficient. Restricting to
$B:u_{1,m_1-1} = -\sum_{i\ge 2} u_{i,m_i-1}$ thus has the effect of
surpressing the term in $du_{1,m_1-1}$ and changing the coefficients
$g_{i,m_i-1}$ by a constant for $i\ge 2$. Expressed in this way, it
follows that the coefficient of $du_{ij}$ in $\Xi|_B$ involves only
monomials in $u_{ip}$ for the same $i$. Giving $u_{ij}$ and $du_{ij}$
both weight $j$, the terms in $g_{ij}du_{ij}$ all have weight $\le
m_i-1$. It
follows from formulas \eqref{3.34}-\eqref{3.36} that, writing $U_{ij} =
u_{ij}-b_{ij}$ so $U_{ij}(b)=0$, we may write
$$\Xi|_B = \sum G_{ij}(U)dU_{ij}.
$$
Here $G_{ij}(0)=0$. giving $U_{ij}$ and $dU_{ij}$ weights $j$, all terms
with first index $i$ have weights $\le m_i-1$. All terms of the form
$$U_{ij}dU_{i,m_i-1-j},\quad 0\le j\le m_i-1,\ i\ge 2\quad(\text{resp.}\
i=1,\ 1\le j\le m_1-2)
$$
occur with nonzero coefficient. Notice that replacing $u_{ij}$ with
$U_{ij}+b_{ij}$ introduces monomials of lower degree, but these have
weight $< m_i-1$.
It follows that $\text{disc}(\Xi|_{\widehat{{\mathcal O}}_{B,b}})$ is the
determinant of a matrix
$$M = \begin{pmatrix}M_1 & 0 &\hdots & \hdots & 0\\
0 & M_2 & 0 & \hdots & 0 \\
\vdots & \vdots & \vdots &\vdots & \vdots \\
0 & 0 & \hdots & \hdots & M_r
\end{pmatrix}
$$
where $M_i$ is symmetric, $m_i \times m_i$ (resp. $(m_1-2) \times
(m_1-2)$), and has the shape
$$\begin{pmatrix}\hdots &\hdots &\hdots &\hdots & \bullet \\
\hdots &\hdots &\hdots &\bullet & 0 \\
\hdots &\hdots &\bullet & 0 & 0 \\
\vdots &\vdots &\vdots &\vdots &\vdots \\
\bullet & 0 & \hdots & \hdots & 0
\end{pmatrix}
$$
with the entries $\bullet$ non-zero.
Mod squares, $\det(M_i)$ is $1$ if $m_i$ is even, and is given by
$$\frac{1}{2}\cdot \text{coefficient
of}(U_{i,\frac{m_i-1}{2}}dU_{i,\frac{m_i-1}{2}})
$$
if $m_i$ is odd. Writing
\eq{\Psi|_B = \sum a_i \frac{du_{i0}}{u_{i0}} + dF
}{ins1}
as just above \eqref{3.49} with $F=\sum a_I u^I$ we find
$$\frac{1}{2} d\log\text{disc}(\Xi|_{\widehat{{\mathcal O}}_{B,b}}) =
\frac{1}{2}\sum\frac{da_I}{a_I}= \frac{1}{2}d\log(\kappa).
$$
(The sum on the right is over all $I$ such that
$u^I = (u_{i,\frac{m_i-1}{2}})^2$.)
\end{proof}
Another interpretation of the $2$-torsion is the following. As in
\eqref{2.27} for $s_i$ a
local parameter at $c_i\in {\mathcal D}$ and $t_i$ another copy of $s_i$ (so
$s_i-t_i$ is a local defining equation for the diagonal in $(C\times C$),
the pullback of
$u_{ip}$ to $K((s_i))$ is the coefficient $s_i^{-(p+1)}$ of $t_i^p$ in
$(s_i-t_i)^{-1}$. It follows from \eqref{3.30} that
\begin{gather} \label{pbtau}
\tau_{ij} \ \text{pulls back
to} \
\frac{-ds_i}{s_i^{j+1}}.
\end{gather}
Write the polar part of the connection at
$s_i=0$ in the form $(g_0+g_1s_i+\ldots)\frac{ds_i}{s_i^{m_i}}$. Since
$\Xi = \sum_{i,j}\lambda_{ij}\tau_{ij}$ pulls back to this connection
form, we get $g_0 = -\lambda_{i,m_i-1}$. On the other hand, again from
\eqref{3.30} the coefficient of
$u_{i,0}^{-2}u_{i,\frac{m_i-1}{2}}du_{i,\frac{m_i-1}{2}}$ in
$\Xi$ is $-\lambda_{i,m_i-1}$ if $m_i$ is odd. This coefficient is
the contribution to ${\rm disc}(\Xi|_{{\mathcal O}_{B,b}})$ from the
point $c_i\in {\mathcal D}$,
so we conclude
\begin{thm} Write the relative connection at $c_i\in {\mathcal D}$ in the form
$(g_{i,0}+g_{i,1}s_i+\ldots)\frac{ds_i}{s_i^{m_i}}$. Then the Gau{\ss}-Manin
connection on
$H^{m-2}_{DR/K}(B,({\mathcal O},d+\Xi))$ is isomorphic to
$$(d+\Psi)|_{\{b\}} +
\sum_i \frac{m_i}{2}d\log(g_{i,0}(0)).
$$
\end{thm}
\begin{defn}\label{def3.16} With notation as above, write
$$\tau(L) = \sum_i \frac{m_i}{2}d\log(g_{i,0}(0)).
$$
\end{defn}
To summarize, we have proven
\begin{thm}\label{thm3.13}Let $C/K$ be a complete smooth curve of
genus $g$ over a field $K\supset k$. Let
$\nabla_{/K}:L \to L\otimes\Omega^1_{C/K}({\mathcal D})$
be a connection,
such that
$$\Big(L \to L\otimes\Omega^1_{C/K}({\mathcal D})\Big) \to
\Big(j_*L|(C-D) \to j_*L\otimes\Omega^1_{C/K}|(C-D)\Big) $$
is a quasiisomorphism. This implies that
the divisor ${\mathcal D}$ is minimal such that $\nabla|_{C-D}$ extends with
values in $\Omega^1_{C/K}({\mathcal D})$ (see section 4, \eqref{4.2}).
We also assume that $L$ has degree 0, and that
the connection on $L|_{C-D}$ lifts
to an integrable, absolute (i.e. $/k$) connection
$\tilde\nabla$.
Then
\eq{\nabla|_{{\mathcal D}}:L|_{\mathcal D} \to L\otimes\Omega^1_{C/K}({\mathcal D})|_{\mathcal D}
}{3.57}
is an ${\mathcal O}_{\mathcal D}$-linear isomorphism and determines a trivialization of
$$\Omega^1_{C/K}({\mathcal D})|_{\mathcal D}$$
Write $J_{\mathcal D}$ for the generalized jacobian and
$J_{{\mathcal D},N}$ for the torseur of divisors of degree $N:=2g-2+\deg
{\mathcal D}$. The above trivialization corresponds to a $K$-point $b\in
J_{{\mathcal D},N}$. Write $\pi_N : (C-D)^N \to J_{{\mathcal D},N}$ for the natural map,
and let $(L_N,\tilde\nabla_N)$ be the evident bundle and absolute
connection on $(C-D)^N$. Then there exists a unique invariant,
absolute connection $({\mathcal L},\Phi)$ on $J_{{\mathcal D},N}$ such that
$\pi_N^*({\mathcal L},\Phi)=(L_N,\tilde\nabla_N)$. Moreover, we have
\eq{\Big(\det(H^*_{DR}(C-D,(L,\nabla)),\nabla_{GM}\Big)^{-1}
\cong ({\mathcal L},\Phi)|_{\{b\}} + \tau(L)
}{3.58}
where $\tau(L)$ is as in definition \ref{def3.16}.
\end{thm}
\begin{remark} \label{rmk3.14}
$2$-torsion also occurs in the determinant of de Rham cohomology for the
trivial connection \cite{BE}. By virtue of the following lemma this can
only happen when the variety has even dimension.
\end{remark}
\begin{lem} \label{lem3.15}
Let $X/K$ be a smooth projective variety of odd dimension $n=2m+1$ over a
function field in characteristic $0$. Then the Gau{\ss}-Manin determinant
$${\rm det}( H_{DR}(X/K), d)$$
is trivial in $\Omega^1_K/d\log K^*$.
\end{lem}
\begin{proof} The strong Lefschetz theorem identifies the determinant
connections on $H_{DR}^p$ and $H_{DR}^{2n-p}$ so we need only consider
the connection on $\det H_{DR}^{n}$. As well known, the Poincar\'e duality
morphism
$$\varphi: H^n_{DR}(X/K) \otimes H^n_{DR}(X/K)
\to H^{2n}_{DR}(X/K)= K$$
is compatible with the Gau{\ss}-Manin connection, which is
trivial on $H^{2n}_{DR}(X/K)= K$.
On the other hand, it is alternating, thus its determinant
$${\rm det} (\varphi): {\rm det}(H^n_{DR}(X/K)) \otimes {\rm
det}(H^n_{DR}(X/K)) \to K$$
fulfills
$$ {\rm det} (\varphi)(e\otimes e)= p^2 \cdot 1$$
where $p\in K^*$ is the Pfaffian of the determinant
of $\varphi$, written in the basis $e$. Thus if
$\nabla(e)=\alpha \otimes e$, one has
$$ {\rm det}(\varphi)(\nabla(e\otimes e)) = 2 \alpha p^2 \cdot
1= 2p d(p) \cdot 1.$$
Thus $\alpha = d\log p$ and the determinant of the
Gau{\ss}-Manin connection is trivial.
\end{proof}
\section{Product and Trace}
In this section, we introduce a product which is reminiscent of
Deligne's product explained in \cite{Dsm}.
We keep the notations of sections 2 and 3
for $C/K,\ (L, \nabla),\ j:U=C-D \to C$, ${\mathcal D}=\sum
m_ic_i$, and ${\mathcal D}' = {\mathcal D} - D$. Further,
\begin{gather}\label{4.1}
\nabla: L \to L\otimes \Omega^1_{U}
\end{gather}
is an absolute connection with vertical curvature $\nabla^2(L)\subset
L\otimes\Omega^2_K\otimes K(X)$. Let $\nabla_{/K}: {\mathcal L} \to {\mathcal L} \otimes
\Omega^1_{C/K}({\mathcal D})$ be an extension of $(L, \nabla_{/K})$ such that
\begin{gather}\label{4.2}
\{{\mathcal L} \to {\mathcal L}
\otimes \Omega^1_{C/K}({\mathcal D})\} \to
\{j_*L \to j_*L\otimes \Omega^1_{U/K}\}
\end{gather}
is a quasiisomorphism. We assume $\nabla_{/K}$ has a pole at every $c_i\in
D$. Note this implies that $\nabla_{/K}$ does not factor through
$\Omega^1_{C/K}({\mathcal D}-c_i)$ for any $i$. Indeed, by assumption, the complex
\eq{j_*L/{\mathcal L} \to (j_*L/{\mathcal L})\otimes\Omega^1_{C/K}({\mathcal D})
}{4.3}
is acyclic. Take $e$ a local basis of ${\mathcal L}$ at $c_i$ and $z$ a local
parameter, and suppose the connection can be written locally as
$\nabla_{/K}e =
a(z)dz/z^{m-1}e$ with $m=m_i$. Then $\nabla_{/K}(z^{-1}e) =
(a(z)dz/z^{m}-z^{-2}dz)e$. The assumption that $\nabla_{/K}$
does have a pole at
$c_i$ implies that $m\ge 2$,
so $z^{-1}e$ would represent a nontrivial element in
$H^0$ of the complex \eqref{4.3}, a contradiction.
By lemma \ref{lem3.1} we know that the verticality condition
implies that the absolute connection extends as
\begin{gather}\label{4.3bis}
\nabla: {\mathcal L} \to {\mathcal L}\otimes
\Omega^1_{C}\langle D \rangle({\mathcal D}').
\end{gather}
{From} now on, we fix such a $({\mathcal L}, \nabla)$.
As we have seen, the map ${\mathcal L}|_{\mathcal D} \to {\mathcal L}\otimes\Omega^1_{C/K}({\mathcal D})|_{\mathcal D}$ is function
linear. Since the connection does not factor through lower order poles, this
gives a trivialization (denoted $\text{triv}(\nabla)$) of $\Omega^1_{C/K}({\mathcal D})|_{\mathcal D}$. We
have
\begin{gather}\label{4.4}
(c_1(\Omega^1_{C/K}({\mathcal D})), {\rm triv \ }(\nabla)) \in
\H^1(C, {\mathcal O}^*_C \to {\mathcal O}^*_{{\mathcal D}}) \\
({\mathcal L}, \nabla) \in \H^1(C, {\mathcal O}^*_C \to \Omega^1_C\langle D \rangle({\mathcal D}')).\notag
\end{gather}
The aim of this section is to define a product
\begin{gather}\label{4.5}
\cup: \H^1(C, {\mathcal O}^*_C \to {\mathcal O}^*_{{\mathcal D}}) \times
\H^1(C, {\mathcal O}^*_C \to \Omega^1_C\langle D \rangle({\mathcal D}')) \notag \\
\to \H^2(C, {\mathcal K}_2 \to \Omega^2_C)
\end{gather}
Here ${\mathcal K}_2$ is the Milnor sheaf associated to $K_2$, and the map ${\mathcal K}_2 \to
\Omega^2_C$ is the $d\log$ map $\{a,b\} \mapsto
\frac{da}{a}\wedge\frac{db}{b}$. For a more detailed study of characteristic
classes for connections defined in the hypercohomology of such complexes, the
reader is referred to \cite{E}. In addition, we will define a trace
\begin{gather}\label{4.6}
{\rm Tr \ }:\H^2(C, {\mathcal K}_2 \to \Omega^2_C) \to \Omega^1_K/d\log K^*
\end{gather}
We write
\eq{A\cdot B := \text{Tr}(A\cup B)
}{4.6bis}
so for example
\begin{multline}\label{4.7}
(c_1(\Omega^1_{C/K}({\mathcal D})), {\rm triv \ }(\nabla))\cdot ({\mathcal L}, \nabla)
:= \\
{\rm Tr \ }((c_1(\Omega^1_{C/K}({\mathcal D})), {\rm triv \ }(\nabla))\cup
({\mathcal L}, \nabla))
\end{multline}
Let
${\mathcal O}^*_{C, {\mathcal D}}= {\rm Ker \ } ({\mathcal O}^*_C \to {\mathcal O}^*_{{\mathcal D}})$.
Then
\begin{lem}\label{lem4.1}
$d \log {\mathcal O}^*_{C, {\mathcal D}} \wedge \Omega^1_C\langle D \rangle({\mathcal D}') \subset
\Omega^2_C$
\end{lem}
\begin{proof}
Since ${\mathcal O}^*_{C, {\mathcal D}}\subset 1+ {\mathcal I}_{{\mathcal D}}$, where
${\mathcal I}_{{\mathcal D}}$ is the ideal sheaf of ${\mathcal D}$,
$$d\log
{\mathcal O}^*_{C, {\mathcal D}} \subset {\mathcal O}_C d {\mathcal I}_{{\mathcal D}} \subset \Omega^1_C(*D).$$
Also one has
$d {\mathcal I}_{{\mathcal D}} \subset {\mathcal I}_{{\mathcal D}}\otimes_{{\mathcal O}_C}\Omega^1_C\langle D \rangle$.
Thus $$
d \log {\mathcal O}^*_{C, {\mathcal D}} \wedge \Omega^1_C\langle D \rangle({\mathcal D}')
\subset {\mathcal I}_D \otimes_{{\mathcal O}_C}\Omega^2_C\langle D \rangle \subset \Omega^2_C.$$
\end{proof}
We define $\cup$ by
\begin{gather}\label{4.8}
{\mathcal O}^*_{C, {\mathcal D}} \cup {\mathcal O}^*_C \to {\mathcal K}_2 \\
(\lambda, c) \mapsto \{\lambda, c\} \notag \\
{\mathcal O}^*_{C, {\mathcal D}} \cup \Omega^1_C\langle D \rangle({\mathcal D}') \to \Omega^2_C \notag \\
(\lambda, \omega) \mapsto d \log \lambda \wedge \omega.\notag
\end{gather}
Concretely, we can write the product in terms of Cech cocyles. Here
${\mathcal C}^i$ refers to Cech cochains, $\delta$ is the Cech coboundary, and $d$ is a
boundary in the complex:
\begin{gather}
(\lambda_{ij}, \mu_i) \in ({\mathcal C}^1({\mathcal O}^*_C) \times
{\mathcal C}^0({\mathcal O}^*_{{\mathcal D}}))_{d-\delta} \notag \\
(c_{ij}, \omega_i) \in ({\mathcal C}^1({\mathcal O}^*_C) \times
{\mathcal C}^0(\Omega^1_C\langle D \rangle({\mathcal D}'))_{d-\delta}\notag
\end{gather}
one has
\begin{gather}\label{4.9}
(\lambda, \mu) \cup (c, \omega) =
(\{\lambda_{ij}, c_{jk}\}, d\log \lambda_{ij} \wedge \omega_j,
-d \log \tilde{\mu_i} \wedge \omega_i ) \\
\in ({\mathcal C}^2({\mathcal K}_2) \times {\mathcal C}^1(\Omega^2_C\langle D \rangle({\mathcal D}')) \times
{\mathcal C}^0(\Omega^2_C\langle D \rangle({\mathcal D}')/\Omega^2_C))_{d+\delta}\notag
\end{gather}
where $\tilde{\mu_i} \in {\mathcal C}^0({\mathcal O}^*_C)$ is a local lifting of
$\mu_i$. Note we have replaced the complex ${\mathcal K}_2 \to \Omega^2_C$ with the
quasiisomorphic complex
$${\mathcal K}_2 \to \Omega^2_C\langle D \rangle({\mathcal D}') \to
\Omega^2_C\langle D \rangle({\mathcal D}')/\Omega^2_C.
$$
\begin{prop}\label{prop4.2}
The product $\cup$ extends to
\begin{gather}
\cup: \H^1(C, {\mathcal O}^*_C \to {\mathcal O}^*_{{\mathcal D}}) \times
\H^1(C, j_*{\mathcal O}^*_U \to \Omega^1_C\langle D \rangle({\mathcal D}')) \notag \\
\to \H^2(C, {\mathcal K}_2 \to \Omega^2_C).\notag
\end{gather}
\end{prop}
\begin{proof}
The map
$$\H^1(C, {\mathcal O}^*_C \to \Omega^1_C\langle D \rangle({\mathcal D}'))
\to \H^1(C, j_*{\mathcal O}^*_U \to \Omega^1_C\langle D \rangle({\mathcal D}'))$$
is surjective,
and its kernel is the ${\mathbb Z}$-module generated by $({\mathcal O}(D_i),
d_i)$ where $d_i$ is the connection with logarithmic poles along
$D_i$ with residue -1.
Let $z_1$ be a local coordinate around $c_1$.
Let $U_i$ be a Cech covering of $C$,
with $c_1 \in U_1 \subset V_1$, and $c_1 \notin U_i, i \neq 1$. Assume $c_1$
is the only zero or pole of $z_1$ on $U_1$. Let
$$(\lambda, \mu) \in \H^1(C,{\mathcal O}^*_C\to{\mathcal O}^*_{{\mathcal D}})
$$
be a Cech representative of a class in $ \H^1(C, {\mathcal O}^*_C \to
{\mathcal O}^*_{{\mathcal D}})$. Then $(c_{ij}, \omega_i)$ with
$c_{1j}= z_1^{-1}$, $c_{ij}= 1$ for $i \neq 1$, $\omega_1=-d
\log z_1$, $\omega_i =0$ for $i\neq 1$ is a Cech representative
of $({\mathcal O}(D_1),d_1)$.
Thus considering $Z \in {\mathcal C}^0({\mathcal O}_C[z_1^{-1}]^*)$ with $Z_1=z_1$ and
$Z_i=1$ for $i\neq 1$,
the cocyle of \eqref{4.6} is just the coboundary
$$(d-\delta)(\{\lambda_{ij}, Z_j \}, d\log \tilde{\mu_i} \wedge d \log
Z_i) \in (d-\delta)({\mathcal C}^1({\mathcal K}_2) \times {\mathcal C}^0(\Omega^2_C\langle D \rangle({\mathcal D}')).$$
(Note $Z_j$ is invertible on $U_{ij}$ for $i\neq j$ so the $K_2$-cochain is
defined.)
\end{proof}
Now we define the trace. We have (with standard $K$-theoretic
notation, \cite{B})
\begin{gather}\label{4.10}
H^2(C, {\mathcal K}_2)=0 \\
{\rm Nm \ } : H^1(C, {\mathcal K}_2) = \{\oplus _{x\in C^{(1)}} \kappa(x)^*\}/
{\rm Tame}(K_2(K(C))) \to K^* \notag\\
\sum_x \varphi_x \mapsto \Pi_x {\rm Nm \ }(\varphi_x) \notag
\end{gather}
and of course $H^1(C, \Omega^2_C)= \Omega^1_K \otimes H^1(C,
\Omega^1_{C/K})= \Omega^1_K$.
This defines
\begin{gather}\label{4.11}
{\rm Tr \ }:\H^2(C, {\mathcal K}_2 \to \Omega^2_C)= H^1(C, \Omega^2_C)/H^1(C,
{\mathcal K}_2)
\to \Omega^1_K/d \log K^* .
\end{gather}
\begin{lem}\label{lem4.3}
The trace
\begin{gather}
{\rm Tr \ }:\H^2(C, {\mathcal K}_2 \to \Omega^2_C)=
\H^2(C, {\mathcal K}_2 \to \Omega^2_C\langle D \rangle({\mathcal D}') \to
\Omega^2_C\langle D \rangle({\mathcal D}')/\Omega^2_C) \notag \\
\to \Omega^1_K/d\log K^* \notag
\end{gather}
factors through
\begin{gather}\label{4.13a}
\H^2\Big(C, {\mathcal K}_2 \to \Omega^1_K \otimes \Omega^1_{C/K}({\mathcal D}) \to
\Omega^1_K \otimes (\Omega^1_{C/K}({\mathcal D})/\Omega^1_{C/K})\Big) \\
\cong \Omega^1_K/d \log K^* \notag \\
\cong \Omega^1_K \otimes_K H^0({\mathcal D}, \omega_{{\mathcal D}/K})/
\H^1(C, {\mathcal K}_2 \to \Omega^1_K \otimes \Omega^1_{C/K}({\mathcal D})) \notag
\end{gather}
where $\omega_{{\mathcal D}/K}$ is the relative dualizing sheaf of the
scheme ${\mathcal D}$, containing $K \cong \omega_{D/K}$.
\end{lem}
\begin{proof}Note that
$$\Omega^1_K\otimes_K \Omega^1_{C/K}({\mathcal D}) \cong
\Omega^2_C\langle D \rangle({\mathcal D}')/(\Omega^2_K\otimes{\mathcal O}_C({\mathcal D}') )
$$
so the complex in \eqref{4.13a} is indeed a quotient. From the diagram
$$\minCDarrowwidth.5cm\begin{CD}@. {\mathcal K}_2 @= {\mathcal K}_2 @. @. \\
@. @VVV @VVV @. \\
0 @>>> \Omega^1_K\otimes\Omega^1_{C/K} @>>> \Omega^1_K\otimes\Omega^1_{C/K}({\mathcal D}) @>>> \Omega^1_K\otimes(\Omega^1_{C/K}({\mathcal D})/\Omega^1_{C/K})
@>>> 0
\end{CD}
$$
one deduces that the left hand side of \eqref{4.13a} is isomorphic to
$$\H^2(C,{\mathcal K}_2 \to \Omega^1_K\otimes\Omega^1_{C/K})\cong \text{coker}(H^1(C,{\mathcal K}_2) \to \Omega^1_K
\otimes H^1(C,\Omega^1_{C/K})).
$$
The right hand side here is identified under the norm with
$\Omega^1_K/d \log K^*$, which proves the second equality. The third
one comes from the map $$\Omega^1_K\otimes \omega_{{\mathcal D}/K}[-2]
\to \{{\mathcal K}_2 \to \Omega^1_K \otimes \Omega^1_{C/K}({\mathcal D}) \to
\Omega^1_K \otimes \omega_{{\mathcal D}/K}\}$$ and the vanishing of
$\H^2({\mathcal K}_2 \Omega^1_K \otimes \Omega^1_C({\mathcal D}))$. Note that this
cumbersome way of writing this cohomology allows to write local
contribution of a class in this cohomology group.
\end{proof}
The first main result of this section is the following
\begin{thm}\label{thm4.4}
Let $({\mathcal L}, \nabla)$ and $({\mathcal L}', \nabla')$ be two extensions of
the vertical connection
$(L, \nabla)$ on $U$ as above satisfying the quasiisomorphism
condition \eqref{4.2}. Then, with notation as in \eqref{4.8},
$$((c_1(\Omega^1_{C/K}), {\rm triv \ } \nabla)\cdot ({\mathcal L},
\nabla))= ((c_1(\Omega^1_{C/K}), {\rm triv \ } \nabla') \cdot ({\mathcal L}',
\nabla')).$$
\end{thm}
\begin{proof}
The quasiisomorphism condition is local about each point of $D$, so we
may assume our line bundles are ${\mathcal L}(\nu c)\subset {\mathcal L}$ for some $\nu
< 0$ and $c\in D$.
Choose local coordinates $z_i$ near $c_i$ and a Cech covering
$U_i$ of $C$ such that $c_i \in U_i$, $z_i \in {\mathcal O}^*(U_i -c_i)$,
$c_i \notin U_j$ for
$i\neq j$. Let us denote by
$$(c_1(\Omega^1_{C/K}), z_i) \in \H^1(C, {\mathcal O}^*_C \to
{\mathcal O}^*_{{\mathcal D}})$$ the class defined by the local trivialization
$$\frac{dz_i}{z_i^{m_i}}: {\mathcal O}_{m_ic_i} \to \Omega^1_{C/K}({\mathcal D})
\otimes {\mathcal O}_{m_ic_i}.$$
Let $({\mathcal L}, \nabla) = (c_{ij}, \omega_i)$. Then
$\omega_i= a_i \frac{dz_i}{z_i^{m_i}} +
\frac{b_i}{z_i^{m_i-1}},$ with $a_i \in {\mathcal O}_C$ such that
$a_i|_{m_ic_i} \in {\mathcal O}^*_{m_ic_i}$ and
$b_i \in \Omega^1_K \otimes {\mathcal O}_C.$ Suppose $c=c_i$. We drop the index
$i$ for convenience. One has
\begin{gather}\label{bigform}
((c_1(\Omega^1_{C/K}), {\rm triv \ } \nabla)\cdot ({\mathcal L},
\nabla))= \\
(c_1(\Omega^1_{C/K}), z)\cdot ({\mathcal L},
\nabla) + \Big(0, 0,- d\log (a) \wedge (a \frac{dz}{z^{m}} +
\frac{b}{z^{m-1}})\Big). \notag
\end{gather}
where the last term is a cocycle as in \eqref{4.10} or the quotient
complex \eqref{4.13}. For $({\mathcal L}(\nu c),\nabla(\nu c))$ one replaces
$a$ by $a-\nu z^{m-1}$,
leaving $b$ and $m$ unchanged.
By proposition \ref{prop4.2}, one has
\begin{gather}
(c_1(\Omega^1_{C/K}), z)\cdot ({\mathcal L},
\nabla)= (c_1(\Omega^1_{C/K}), z)\cdot ({\mathcal L}(\nu c),
\nabla(\nu c)) \notag
\end{gather}
Thus
\begin{gather}
((c_1(\Omega^1_{C/K}), {\rm triv \ } \nabla)\cdot ({\mathcal L},
\nabla)) -((c_1(\Omega^1_{C/K}), {\rm triv \ } \nabla') \cdot ({\mathcal L}',
\nabla')) \notag \\
= \Big(0,0, (d(a- \nu z^{m-1}) - d(a))\wedge \frac{dz}{z^m} + d\log
(\frac{a - \nu z^{m-1}}{a}) \wedge \frac{b}{z^{m-1}}\Big)
\notag \\
= \Big(0,0,\frac{ \nu (m-1) d\log z \wedge \frac{b}{a}}{(1-\nu
\frac{z^{m-1}}{a})}\Big).\notag
\end{gather}
The nontrivial part in the last expression is computed in
$$H^0(\Omega^2_C\langle c\rangle/\Omega^2_C)\cong \Omega^1_c.
$$
Computing using the residue at $c$ we find the above difference is
$$\Big(0,0,\nu (m-1)\frac{b(c)}{a(c)}\Big)
$$
The verticality condition for the curvature reads
$$
$$
$$ da \wedge \frac{dz}{z^m} + \frac{db}{z^{m-1}}
-(m-1)\frac{dz}{z^m} \wedge b =0\in \Omega^1_K \otimes
\Omega^1_{C/K}({\mathcal D}).
$$
In particular, $(da - (m-1) b)_{|c} = 0$.
The difference of the two products
is therefore $(0,0, d\log a^\nu)$, which vanishes in $\Omega^1_K /d\log K^*$.
\end{proof}
\begin{remark}A version of the formula \eqref{bigform} in higher rank plays a
central role in section 5.
\end{remark}
Suppose now
\begin{gather}\label{4.12}
m_i=1 {\rm \ for \ all \ } i.
\end{gather}
In this case, the class $
(c_1(\Omega^1_{C/K}(D), z_i) \in \H^1(C, {\mathcal O}^*_C \to {\mathcal O}^*_{{\mathcal D}})$
as defined in the proof of theorem \ref{thm4.4} does not in fact depend
on the choice of the local coordinate $z_i$. Indeed, the
trivialization $${\mathcal O}_D \to (\Omega^1_{C/K}(D)/\Omega^1_{C/K}=
{\mathcal O}_D), 1 \mapsto \frac{dz_i}{z_i}$$ is just the canonical
identification given by the residue along $c_i$. In other words,
the class $(c_1(\Omega^1_{C/K}(D), z_i)$ is what is denoted by
$(c_1(\Omega^1_{C/K}(D), {\rm res }_D)$
in \cite{BE}, and appears on the right hand side of the
Riemann-Roch formula.
The second main result of this section is
\begin{thm}\label{thm4.5}
Let $({\mathcal L}, \nabla)$ be as above, with $m_i=1$ for
all $i$. Then
\begin{multline*}{\rm det \ } \Big(H^*_{DR}(U, L)),\ \text{Gau\ss\ - Manin
connection}\Big) = \\
- c_1\Big(\Omega^1_{C/K}(D), {\rm
triv }\nabla \Big)\cdot ({\mathcal L}, \nabla).
\end{multline*}
\end{thm}
\begin{proof}
Given the main result of \cite{BE}, and lemma \ref{lem3.15},
the theorem is of course
equivalent to
\begin{gather}\label{4.13}
c_1(\Omega^1_{C/K}(D), {\rm
res}_D) \cdot ({\mathcal L}, \nabla) = c_1(\Omega^1_{C/K}(D), {\rm triv
\ } \nabla) \cdot ({\mathcal L}, \nabla).
\end{gather}
Keeping the same notations as in the proof of theorem
\ref{thm4.4}, one has
\begin{gather}
c_1(\Omega^1_{C/K}(D), {\rm
res}_D) - c_1(\Omega^1_{C/K}(D), {\rm
triv \ } \nabla)= (0, a_i) \notag
\end{gather}
and thus
\begin{gather}
(c_1(\Omega^1_{C/K}(D), {\rm res}_D) - c_1(\Omega^1_{C/K}(D), {\rm
triv \ } \nabla)) \cdot ({\mathcal L}, \nabla) = \notag \\
(0,0, -d (a_i) \wedge d\log z_i - d\log a_i \wedge b_i) = \notag
\\ (0,0, -d(a_i) \wedge d\log z_i). \notag
\end{gather}
This lies in $\Omega^1_K \otimes \omega_{D/K}=\Omega^1_K$ and by
lemma \ref{lem4.3}, its trace factors through $\Omega^1_K
\otimes H^1(C, \Omega^1_{C/K})$. But the image of $\gamma=\sum_i a_i
d\log z_i \in \omega_{D/K}$ in $H^1(C, \Omega^1_{C/K})$ is the
relative Atiyah class
${\rm at }_{/K}({\mathcal L})$ (\cite{EV}, appendix B), thus the image of
$d\gamma= \sum_i d(a_i)\wedge
d\log z_i \in \Omega^1_K \otimes \omega_{D/K}$
in $\Omega^1_K \otimes H^1(C, \Omega^1_{C/K})$
is $d({\rm at }({\mathcal L}))$, where ${\rm at}({\mathcal L}) \in H^1(C,
\Omega^1_C)$ is the absolute Atiyah class of ${\mathcal L}$. Indeed,
$d: H^1(C, \Omega^1_C) \to \Omega^1_K \otimes H^1(C,
\Omega^1_{C/K})$ factors through $H^1(C, \Omega^1_{C/K})$ by
Hodge theory. On the other hand, if $c_{ij} \in {\mathcal C}^1({\mathcal O}^*_C)$
is a cocyle for ${\mathcal L}$, then $d\log c_{ij} \in {\mathcal C}^1(\Omega^1_C)$
is a cocyle for ${\rm at}({\mathcal L})$, and consequently, $d({\rm at }({\mathcal L}))=0$.
\end{proof}
We want to explain briefly a fundamental compatibility satisfied by the pairing
\eqref{4.6bis}. Let $b = \sum b_i$ be a $0$-cycle on $C$ with support disjoint
from ${\mathcal D}$, and let $\nabla:{\mathcal L}\to
{\mathcal L}\otimes\Omega^1_{C}<D>({\mathcal D}')$ be an absolute,
integrable connection. We can interpret $\nabla|_{b_i} \in
\Omega^1_{K(b_i)}/d\log K(b_i)^*$.
\begin{prop}\label{prop4.6} With notation as above, let $[b]\in
\H^1(C,{\mathcal O}_C^* \to {\mathcal O}_{\mathcal D}^*)$ be the class of the $0$-cycle $b$. Then
$$[b]\cdot ({\mathcal L},\nabla) = \sum_i
{\rm Tr}_{K(b_i)/K}(\nabla|_{b_i})\in \Omega^1_K/d\log K^*.
$$
Let $({\mathcal L}_0, \nabla_0)$ be the invariant connection on $J_{\mathbb D}$
which pulls back to $({\mathcal L}, \nabla)$ via the cycle map $i: (C-D)
\to J_{\mathcal D}$ (proposition \ref{prop2.16}). Then
$$[b]\cdot ({\mathcal L},\nabla)= {\rm Tr}_{i_0(b)/} ({\mathcal L}_0, \nabla_0)|_{i_0(b)}.
$$
\end{prop}
\begin{proof}
One reduces easily to the case $b$ is a single $K$-point. Let
$U_2$ be a Zariski-open set containing $D$ and $b$. Shrink $U_2$
if necessary so
there exists $z\in H^0({\mathcal O}_{U_2})$ with $z|_{\mathcal D} = 1$ and $(z)=b$. Let
$U_1=C-\{b\}-{\mathcal D}$ so $C=U_1\cup U_2$. Shrinking the $U_i$ if necessary, we can
assume ${\mathcal L}|_{U_i}\cong {\mathcal O}_{U_i}$, so $({\mathcal L},\nabla)$ is represented by some
cocycle $(\mu_{12},\omega_1,\omega_2)$. Then
$$\nabla|_b = \omega_2|_b \in \Omega^1_K/d\log K^*
$$
On the other hand, by the definition \eqref{4.10},
$[b]\cdot ({\mathcal L},\nabla)$ is represented by the image of the cocycle
$$d\log(z)\wedge \omega_2|_{U_{12}}
\in H^1(C,\Omega^2_C)
\to
\Omega^1_K\otimes_K H^1(C,\Omega^1_{C/K}) \cong \Omega^1_K.
$$
Write $\omega_2 = \omega_2(b)+z\eta_2$ with $\eta_2$ regular on $U_2$. Since
$d\log(z)\wedge z\eta_2$ extends to $U_2$, it is homologous to zero, so
$$[b]\cdot ({\mathcal L},\nabla) = \omega_2(b)[b]\in \Omega^1_K\otimes\H^1(C,\Omega^1_{C/K}) \mapsto
\omega_2(b)\in \Omega^1_K.
$$
Now, since one obviously has
$${\rm Tr} (\nabla|_{b_i}) = {\rm Tr} (\nabla_0)|_{i_0(b_i)}
$$
and the translation $\nabla_0$ is invariant, the second
equality is a direct interpretation of the first one.
\end{proof}
Now we can formulate and prove a variant of theorem \ref{thm3.13}.
\begin{thm}\label{thm4.7}
Let $(C/K, U/K, (L, \nabla),{\mathcal D})$ be as in \eqref{4.1},
\eqref{4.2}, \eqref{4.3}. Then
\begin{gather}\label{4.14}
{\rm det \ }(H^*_{DR}(U, L))= - (c_1(\Omega^1_{C/K}({\mathcal D}), {\rm
triv \ }(\nabla))\cdot ({\mathcal L}, \nabla) \in \Omega^1_K/d\log K^*
\end{gather}
modulo torsion (see remark \ref{rmk3.14}),
where $\nabla_{/K}: {\mathcal L} \to {\mathcal L} \otimes \Omega^1_{C/K}({\mathcal D})$ is
any extension of $(L, \nabla_{/K})$ having poles along all points of
${\mathcal D}$ such that
\begin{gather}
\{{\mathcal L} \to {\mathcal L} \otimes \Omega^1_{C/K}({\mathcal D}) \} \to \{j_*L \to
j_*(L\otimes \Omega^1_{U/K})\}\notag
\end{gather}
is a quasiisomorphism.
\end{thm}
\begin{proof}
If all $m_i=1$ this is just theorem \ref{thm4.5}. So we assume
that $m_1 \geq 2$ in the sequel. Then as in the proof of theorem
\ref{thm4.4}, replacing ${\mathcal L}$ by ${\mathcal L}(-c_1)$ changes $a_1$ to
$(a_1 +z_1^{m_1-1})$ and keeps the rest unchanged. Thus
the quotient complex
$${\mathcal L}/{\mathcal L}(-c_1) \to {\mathcal L}/{\mathcal L}(-c_1) \otimes \Omega^1_{C/K}({\mathcal D})$$
is ${\mathcal O}_{c_1}$-linear and the map is the multiplication by $a_1
\in {\mathcal O}^*_{c_1}$. In particular, ${\mathcal L}(\nu c_1)$ fulfills \eqref{4.2} for
all $\nu \in {\mathbb Z}$ and taking $\nu=- {\rm deg \ } {\mathcal L}$, we may
assume by theorem \ref{thm4.4} that ${\rm deg \ } {\mathcal L}= 0$.
If $H^0_{DR}(U,L) \neq 0$, then there is a meromorphic section $\varphi$ of
${\mathcal L}$ verifying the flatness condition
$$d\varphi + \omega \varphi=0.$$
This implies in particular that $\omega$ has at most logarithmic
poles along $D$, which contradicts the condition $m_1 \geq 2$.
On the other hand, $H^2_{DR}(U, L) =0$ for dimension reasons,
thus we can apply theorem
\ref{thm3.13} together with
proposition \ref{prop4.6} to obtain the result, after we
have replaced $({\mathcal L}, \nabla)$ by $({\mathcal L}, \nabla) \otimes f^*\Big(({\mathcal L},
\nabla)|_{c_0}^{-1}\Big)$ to trivialize the
connection at $c_0$ and applied the projection formula
to this tensor connection.
\end{proof}
We finish this section with an example. Let $U={\mathbb A}^1_K$ be the
affine line over ${\rm Spec}K$, with parameter $t$,
and $\nabla$ be a connection on the trivial bundle. Then
up to a twist by a form of the base, $\nabla$ has equation
$A= df$, where $f= \sum_{i=1}^{m-1}a_i t^i$, $a_i \in K, a_{m-1}
\neq 0$. Write $df= d_Kf + f'dt$ with $d_kf= \sum_{i=1}^{m-2}
da_i t^i$ and $f'= \sum_{i=1}^{m-2} ia_i t^{i-1}$. Let $b_i,
i=1,\ldots, (m-2)$ be the zeroes of $f'$ (defined over $K$ after
some finite field extension), and let $N_\ell(\underline{b})=
\sum_{i=1}^{m-2} b_i^\ell$ be the Newton classes of the zeroes
of $f'$, which of course are expressable in the $a_i$ already on
${\rm Spec}K$. Then the main theorem says
$${\rm det}(GM)^{-1}= \sum_{i=1}^{m-2} da_i N_i(\underline{b}).$$
\section{A Formula in Higher Rank}
In this section, we want to define a sort of non-commutative
product of a higher rank connection with the Chern class of
the dualizing sheaf of $C$ with poles. The notations are as in
the whole article: $C$ is a curve defined over a function field
$K$ over an algebraically closed field $k$ of characteristic 0,
$U$ is an open set such that $D=X-U=\sum_i c_i$ consists of $K-$rational
points. Let $({\mathcal E}, \nabla)$ be a rank $r$-connection on $U$ with vertical
curvature \eqref{2.60}. Let $m_i$ be the multiplicity of the relative
connection at the point $c_i$, that is, the minimal multiplicity
such $\nabla $ factors
$$\nabla_{/K}: E \to E\otimes \Omega^1_{C/K}(\sum_i m_ic_i).$$
Lemma \ref{lem3.1} no longer holds
true in the higher rank case.
We say that the poles of the global connection {\it behave
well} if
\begin{gather} \label{ass5.1}
\nabla: E \to E \otimes \Omega^1_C<D>({\mathcal D}')
\end{gather}
where ${\mathcal D} = \sum_i m_i c_i$ and ${\mathcal D}'= {\mathcal D} - D$.
Let $s=\{s_i\}$ be a trivialization of $\omega_{C/K}({\mathcal D})|_{\mathcal D}
\cong \omega_{{\mathcal D}/K}$. That
is, $s_i\in \omega_{m_ic_i}$ and the map $1\mapsto s_i$ is an isomorphism
${\mathcal O}_{\mathcal D} \cong \omega_{m_ic_i}$. For example, if $z_i$ is a local parameter,
one can take $s_i = \frac{dz_i}{z_i^{m_i}}$. We will abuse notation and write
$s_i$ also for a lifting of the trivialization to a local section of
$\Omega^1_C< D>({\mathcal D}')$. The local matrix of the connection has the shape
\begin{gather}
A_i = g_i s_i + \frac{\eta_i}{z_i^{m_i-1}}
\end{gather}
where $g_i$ and $\eta_i$ are $r\times r$ matrices with
coefficients in ${\mathcal O}_C$ and $\Omega^1_K \otimes {\mathcal O}_C$
respectively. Note that the matrix of functions $g_i$ depends only on the
lifting of $s_i$ to a section of $\omega_{C/K}({\mathcal D})$. (Indeed, the relative
connection has matrix $gs_i$.)
We assume
\begin{gather}
{\rm Image}(g_i) \in M(r \times r, {\mathcal O}_{{\mathcal D}}) \notag \\
\label{ass5.3}
{\rm lies \ in \ } GL(r, {\mathcal O}_{{\mathcal D}})
\end{gather}
Under this assumption, we define
\begin{gather}\label{def5.4}
\{c_1(\omega({\mathcal D})), \nabla\}:= \\
c_1(\omega({\mathcal D}),
s_i) \cdot {\rm det}( \nabla) - \sum_i{\rm res \
Tr}(dg_i g_i^{-1}A_i) \notag \\
\in \Omega^1_K/d\log K^* \notag
\end{gather}
\begin{conj} \label{conj} Assuming \eqref{ass5.1} and
\eqref{ass5.3}, we have
$${\rm det}H^*_{DR}(U, ({\mathcal E}, \nabla))^{-1} =
\{c_1(\omega({\mathcal D})), \nabla\} \in (\Omega^1_K/d\log K^*) \otimes_{{\mathbb Z}} {\mathbb Q}.
$$
\end{conj}
We discuss the assumption \eqref{ass5.3} (see proposition
\ref{prop5.6}) at the end of this section.
The assumption \eqref{ass5.1} on the poles behaving
well is not very well understood. It reflects a sort of
stability in all possible directions for
the poles of the global connection.
First, we justify
the conjecture by establishing some rather surprising invariance
properties for $\{c_1(\omega({\mathcal D})), \nabla\}$.
\begin{lem}\label{lem5.2}Fix an index $i$ and write the connection matrix
locally in the form
$$A = g\frac{dz}{z^m}+\frac{\eta}{z^{m-1}}
$$
where $g$ is an invertible matrix of functions and $\eta$ is a matrix with
entries from $\Omega^1_K\otimes{\mathcal O}_C$. Then \begin{enumerate}\item $\text{res
Tr}(dgg^{-1}A)= \text{res Tr}(dgg^{-1}\frac{\eta}{z^{m-1}})$.
\item $[\eta,g]z^{1-m}$ has no pole at the point $z=0$.
\end{enumerate}
\end{lem}
\begin{proof}The assumption that the curvature is vertical implies
\begin{equation} \label{eq5.5}
dA = dg\frac{dz}{z^m}+d(\frac{\eta}{z^{m-1}}) \equiv A^2 =
[\eta,g]\frac{dz}{z^{2m-1}} \mod\ \Omega^2_K\otimes{\mathcal O}_C[z^{-1}]
\end{equation}
Multiplying through by $z^m$ and contracting against $\frac{\partial}{\partial
z}$ we deduce 2.
For 1, we must show $\text{res Tr}(dg\frac{dz}{z^m}) = 0$. From \ref{eq5.5},
using $\text{Tr}[g,\eta] = 0$ we reduce to showing $\text{res
Tr}d(\frac{\eta}{z^{m-1}}) = 0$.
Since $\eta$ has entries $\Omega^1_K$, one has
$$\text{res
Tr}d(\frac{\eta}{z^{m-1}}) = \text{res
Tr}d_{C/K}(\frac{\eta}{z^{m-1}}).$$
And the residue of an exact form is vanishing.
\end{proof}
\begin{lem}$\{c_1(\omega({\mathcal D})), \nabla\}$ is independent of the choice of the
trivializations $s_i$.
\end{lem}
\begin{proof}First we show independence of the choice of lifting of $s$. As
remarked above,
$g$ is determined by the local lifting of $s$ to $\omega({\mathcal D})$, so
$\{c_1(\omega({\mathcal D})), \nabla\}$ depends only on that choice. If $s$ and $s'$ are
two such local liftings, with $gs=g's'$, we have
$$dgg^{-1}=dg'g'{}^{-1}+d\log(\frac{s'}{s})I= dg'g'{}^{-1}+z^mh
$$
for some $h\in M_r({\mathcal O}_C)$. It follows immediately that
$$\text{res Tr}(dgg^{-1}A) =
\text{res Tr}(dg'g'{}^{-1}A)
$$
as desired.
Next we show independence of the choice of trivializations themselves. Let
$f$ be a rational function on $C$ whose divisor $(f)$ is disjoint from the
singular locus of $\nabla$. It will suffice to show that $s$ and $fs$ as
trivializations give rise to the same invariant, i.e.
\begin{multline}\label{5.5}c_1(\omega({\mathcal D}),
s) \cdot {\rm det}( \nabla) - \sum_i{\rm res \
Tr}(dg_i g_i^{-1}A_i) = \\
c_1(\omega({\mathcal D}),
fs) \cdot {\rm det}( \nabla) - \sum_i{\rm res \
Tr}((dg_i g_i^{-1}-dff^{-1}I)A_i)
\end{multline}
Recall we can calculate $c_1(\omega({\mathcal D}), s) \cdot {\rm det}( \nabla)$ by
choosing $\delta$ a divisor in the linear series $\omega({\mathcal D})$ compatible with
the rigidification $s$ and then restricting $\nabla|_\delta$ and taking the
norm to ${\rm Spec \,}(K)$. Associated to the trivialization $fs$ we may take the
divisor $\delta+(f)$. It follows that
$$c_1(\omega({\mathcal D}),
fs) \cdot {\rm det}( \nabla) - c_1(\omega({\mathcal D}),
s) \cdot {\rm det}( \nabla) = \text{Norm}\det\nabla|_{(f)}
$$
(To get this relation, one could have taken the formula
\eqref{bigform} as well).
On the other hand,
since the formula depends only on the local behavior of $f$
near ${\mathcal D}$, by
suitably choosing $f$, we may assume $\nabla$ is defined by
$A$ in a neighborhood of $(f)$ and that $f\equiv 1$ modulo some large power of
the maximal ideal at the finite set of points where the connection is not
given by $A$. We can interpret $\text{Tr}
(dff^{-1}A)\in \Omega^1_K \otimes \omega_{k(C)}$, so
the sum of the residues over all closed points of $C$ will vanish in $\Omega^1_K$.
Thus
$$\sum_i\text{res Tr}(dff^{-1}A_i) = -\sum_{(f)}\text{res Tr}(dff^{-1}A) =
\text{Tr}(A|_{(f)})
$$
Since the connection matrix for the determinant connection is the trace of the
connection matrix, the contributions to \eqref{5.5} cancel.
\end{proof}
Now consider what happens to the expression
\begin{equation}\{c_1(\omega({\mathcal D})), \nabla\}
\end{equation}
under a change of coordinates given by a matrix $h$ of functions. We have
\begin{equation}A \mapsto A' := hAh^{-1}+dhh^{-1}=
g'\frac{dz}{z^m}+\frac{\eta'}{z^{m-1}}
\end{equation}
Note that $h$ is regular, so $dhh^{-1}$ does not contribute to the polar part of the
connection, i.e.
\begin{equation}g'=hgh^{-1}+z^ma;\quad \eta' = h\eta h^{-1}+z^{m-1}b
\end{equation}
with $a$ and $b$ regular.
We compute
\begin{multline}dg'g'{}^{-1}= d(hgh^{-1})hg^{-1}h^{-1}+ez^m+fz^{m-1}dz = \\
= dhh^{-1}+hdgg^{-1}h^{-1}-hgh^{-1}dhg^{-1}h^{-1}+ez^m+fz^{m-1}dz
\end{multline}
with $e$ and $f$ regular. Thus
\begin{multline}\label{5.11} \text{res Tr}(dg'g'{}^{-1}A') = \text{res
Tr}(dg'g'{}^{-1}hAh^{-1}) = \\ \text{res Tr}(h^{-1}dh A) +
\text{res Tr}(dgg^{-1}A)- \text{res
Tr}(h^{-1}dhg^{-1}Ag).
\end{multline}
Note
\begin{equation}A-g^{-1}Ag = z^{1-m}(\eta - g^{-1}\eta g) =
z^{1-m}g^{-1}(g\eta - \eta g)
\end{equation}
{From} lemma \ref{lem5.2},2, this expression is regular. Plugging into
\eqref{5.11}, we conclude.
\begin{lem}$\{c_1(\omega({\mathcal D})),\nabla\}$ is independent of the choice of basis
for the bundle $E$.
\end{lem}
\begin{remarks}\begin{enumerate}\item The definition of
$\{c_1(\omega({\mathcal D})),\nabla\}$ was inspired by the calculations in section 4
(cf. formula \eqref{bigform}) for a rank one connection. The formula
\begin{gather}
{\rm det}H^*_{DR}(U, ({\mathcal E}, \nabla))^{-1} =
\{c_1(\omega({\mathcal D})), \nabla\}
\end{gather}
when applied to the rank 1 case, gives back the
main theorem of this article.
\item When $m_i=1$ for all $c_i$, that
is when $\nabla$ has regular singularities, then
the argument from theorem \ref{thm4.5} (slighlty modified in the
higher rank case) gives
\begin{gather}
\{c_1(\omega(D)), \nabla\} =
c_1(\omega(D), {\rm res}_D) \cdot {\rm det}(\nabla)
\end{gather}
where, as in theorem \ref{thm4.5}, ${\rm res}_D$ refers to the
natural trivialization coming from the residue.
\item Finally, twisting $\nabla$ by $f^*\alpha$, where $\alpha \in
\Omega^1_K$ comes from the base, changes the right hand side of
the formula by
$$(2g-2 + n -\sum_i m_i)r\alpha = -\chi(H_{DR}^*(U,
({\mathcal E}, \nabla_{/K})))\alpha ,$$
as it should. Here $r={\rm rank} E$.
\end{enumerate}
\end{remarks}
\begin{prop} \label{prop5.6}
Let ${\mathcal D}\subset C$ be a divisor on a smooth curve, and let $f\in {\mathcal O}_{C,{\mathcal D}}$
be a local defining equation for ${\mathcal D}$. Let
$$\nabla : E \to E\otimes\omega({\mathcal D})
$$
be a connection, and write
$$g=\nabla|_{\mathcal D} : E/E(-{\mathcal D}) \to (E({\mathcal D})/E)\otimes\omega
$$
Let $j:E-{\mathcal D} \hookrightarrow E$ be the inclusion and consider the connection
$$j_*j^*\nabla : j_*j^*E \to j_*j^*E\otimes \omega.
$$
There is a natural map
$$\iota:\{E\to E\otimes\omega\} \to \{j_*j^*\nabla : j_*j^*E \to
j_*j^*E\otimes \omega\}.
$$
The map $\iota$ is a quasi-isomorphism if and only if for any
$n\ge 0$ the natural map
$$g-n\cdot id\otimes\frac{df}{f}:E(n{\mathcal D})/E((n-1){\mathcal D}) \to
(E((n+1){\mathcal D})/E(n{\mathcal D}))\otimes\omega
$$
given by $f^{-n}e\mapsto f^{-n}g(e)-nf^{-n-1}e\otimes df$ is a
quasi-isomorphism.
In particular, if,
every point of ${\mathcal D}$ has multiplicity $\ge 2$, then $g$ is an
isomorphism if and only if $\iota$ is a quasi-isomorphism.
\end{prop}
\begin{proof} The usual exact sequence reduces us to showing the
condition is equivalent to
$$j_*j^*E/E \stackrel{\cong}{\to}(j_*j^*E/E)\otimes \omega({\mathcal D}).
$$
Writing as usual ${\mathcal D}={\mathcal D}'+D$ where $D$ is the reduced divisor
with support equal to the
support of ${\mathcal D}$, we have a commutative square
$$\begin{CD} E/E(-{\mathcal D}) @>g -n id\otimes \frac{df}{f} >> (E({\mathcal D}')/E(-D))\otimes
\omega(D) \\ @V\cong V``\cdot f^{-n}" V @V\cong V ``\cdot f^{-n}" V \\
E(n{\mathcal D})/E((n-1){\mathcal D}) @>g>>
(E(n{\mathcal D}+{\mathcal D}')/E((n-1){\mathcal D}+{\mathcal D}'))\! \otimes\! \omega(D)
\end{CD}
$$
The map $\iota$ is a quasi-isomorphism if and only if for all
$n\ge 0$ the map $g$ on the
bottom line of the above square is an isomorphism, and this will
hold only if the top line
is.
In particular, since $\frac{df}{f}$ has poles of order 1,
if all multiplicities are
$\geq 2$, then $\iota$ is a quasiisomorphism if and only if $g$
is an isomorphism.
\end{proof}
We close those remarks by a numerical computation
for $E= \oplus_1^r {\mathcal O}$ on
$U={\mathbb A}^1_K$, with parameter $t$ .
There is only one singular point at $\infty$.
Let us write
$$A = \sum_{i=0}^{m-1}B_it^i + \sum_{i=0}^{m-2} C_i t^idt$$
where the $B_i$ and the $C_i$
are matrices with coefficients in $\Omega^1_K$
respectively $K$. The assumption \ref{ass5.3} means that
$C_{m-2} \in GL(r, K)$. We consider the cases $m=2,3$. For $m=2$ both sides
of the formula are 0, and for $m=3$ they are equal to
\begin{gather}
{\rm Tr}(B_0 -B_1 C_0C_1^{-1} + B_2 C_0C_1^{-1}C_0C_1^{-1})
\end{gather}
Note that $-C_0C_1^{-1}$ is the zero of the ``polynomial''
$C_0 + C_1t$, where $t$ is a ``variable'' of matrices, and thus
the formula could also be written as
\begin{gather}
{\rm Tr} A|_{{\rm zero}(C_0 + C_1t)}
\end{gather}
if it made sense.
For higher $m$, the right hand side should be a sort of restriction of
$\nabla$ to the ``Newton'' classes of $\sum_{i=0}^{m-2} C_i t^i$.
In the remaining part of this section, we show that the product
$\{c_1(\omega({\mathcal D})), \nabla\}$ is a particular case of a more
general product between higher rank connections and a larger
class of trivializations along ${\mathcal D}$.
We consider the tuples $\{E, L, \nabla, {\mathcal D}, g\}$, where
\begin{gather}
\nabla: E \to E \otimes \Omega^1_C<D>({\mathcal D}')
\end{gather}
is a connection on a rank $r$ vector bundle $E$. We denote
by $\nabla|_{{\mathcal D}}$ the ${\mathcal O}_C$-linear map
\begin{gather}
\nabla: E \to E \otimes \Big(\Omega^1_C<D>({\mathcal D}')/\Omega^1_C\Big).
\end{gather}
Further, $L$ is a rank 1
bundle, $g$ is a trivialization
\begin{gather}
g: E|_{{\mathcal D}} \cong E\otimes L|_{{\mathcal D}},
\end{gather}
which fulfills
\begin{gather} \label{comm}
0=[g, \nabla|_{{\mathcal D}}]: E|_{{\mathcal D}} \to E\otimes L|_{{\mathcal D}}
\end{gather}
By lemma \ref{lem5.2}, if $L = \omega_C({\mathcal D})$ and $g$ is the
trivialization $E|_{{\mathcal D}} \to E\otimes \Big(\omega_C({\mathcal D})/\omega_C\Big)$
arising from the principal part of $\nabla_{C/K}$, then the
condition \ref{comm} is fulfilled.
Let us introduce the cocyles of the tuple $\{E,L, \nabla, {\mathcal D},
g\}$, as in section 4. If $E$ has cocycle $c_{ij} \in GL(r,
{\mathcal O}_C)$, $L$ has cocycle $\lambda_{ij} \in {\mathcal O}^*_C$, $g$ has
cocyle $\mu_i \in GL(r, {\mathcal O}_{{\mathcal D}})$, and $\nabla$ has cocyle
$\omega_i \in M(r \times r, \Omega^1_C<D>({\mathcal D}'))$ then one has
\begin{gather} \label{coc}
dc_{ij}c_{ij}^{-1} =
\omega_i-c_{ij}\omega_jc_{ij}^{-1} \\
\mu_i=c_{ij}\mu_jc_{ij}^{-1}\lambda_{ij}.
\end{gather}
The commutativity condition \ref{comm} then reads
\begin{gather}\label{comm2}
[\mu_i, \omega_i] = 0.
\end{gather}
Let $\tilde{\mu}_i \in GL(r, {\mathcal O}_C)$ be a local lifting ot
$\mu_i\in GL(r, {\mathcal O}_{{\mathcal D}})$. Then one has
\begin{thm}\label{prod}
The cochain
\begin{equation} \label{coch}
\Big\lbrace\{\lambda_{ij},\det(c_{jk})\},
d\log\lambda_{ij}\wedge {\rm Tr}(\omega_j),
{\rm Tr}(-d\tilde\mu_i\tilde\mu_i^{-1}\wedge
\omega_i)\Big\rbrace
\end{equation}
is a cocyle in
\begin{equation}{\mathcal K}_2 \to {\Omega^2_X\langle D\rangle({\mathcal D}')}\to \Big({\Omega^2_X\langle D\rangle({\mathcal D}')}/{\Omega^2_X}\Big)
\end{equation}
and defines a cohomology class $\{c_1(L), g, \nabla\}$ in
\begin{equation}\H^2\Big(C,{\mathcal K}_2 \to {\Omega^2_X\langle D\rangle({\mathcal D}')}\to \Big({\Omega^2_X\langle D\rangle({\mathcal D}')}/{\Omega^2_X}\Big)\Big)
\end{equation}
\end{thm}
\begin{proof}
First
\begin{multline}\delta(d\log(\lambda)\wedge {\rm Tr}(\omega))(ijk)= \\
d\log(\lambda_{ij})\wedge {\rm Tr}(\omega_j)+
d\log(\lambda_{jk})\wedge {\rm Tr}(\omega_k) -
d\log(\lambda_{ik})\wedge {\rm Tr}(\omega_k) = \\
d\log(\lambda_{ik})\wedge {\rm Tr}(\omega_j-\omega_k) - d\log(\lambda_{jk})\wedge
{\rm Tr}(-\omega_k+\omega_j) = \\
d\log(\lambda_{ij})\wedge {\rm Tr}(\omega_j-\omega_k) = d\log(\lambda_{ij})\wedge
d\log(\det(c_{jk})).
\end{multline}
Next computing mod the ideal of ${\mathcal D}$ and so ignoring the tilde,
\begin{multline} \delta({\rm Tr}(-d\tilde\mu_i\tilde\mu_i^{-1}\wedge
\omega_i))(ij) = {\rm Tr}(-d\tilde\mu_j\tilde\mu_j^{-1}\wedge
\omega_j + d\tilde\mu_i\tilde\mu_i^{-1}\wedge
\omega_i) = \\
{\rm Tr}\Big((d\lambda_{ij}\lambda_{ij}^{-1}\cdot I +c_{ij}^{-1}dc_{ij}
-c_{ij}^{-1}d\mu_i\mu_i^{-1}c_{ij}-c_{ij}^{-1}\mu_idc_{ij}c_{ij}^{-1}
\mu_i^{-1}c_{ij}) \wedge\omega_j+ \\
+d\mu_i\mu_i^{-1}\wedge\omega_i\Big) =
{\rm Tr}\Big((d\lambda_{ij}\lambda_{ij}^{-1}\cdot I + c_{ij}^{-1}dc_{ij}
-c_{ij}^{-1}d\mu_i\mu_i^{-1}c_{ij}\\ -c_{ij}^{-1}\mu_idc_{ij}c_{ij}^{-1}
\mu_i^{-1}c_{ij}) \wedge(c_{ij}^{-1}\omega_ic_{ij}-c_{ij}^{-1}dc_{ij})+
d\mu_i\mu_i^{-1}\wedge\omega_i\Big)\\
\end{multline}
By our commutation assumption \ref{comm}
\begin{equation}{\rm Tr}(c_{ij}^{-1}dc_{ij}\wedge
c_{ij}^{-1}\omega_ic_{ij}) =
{\rm Tr}(c_{ij}^{-1}\mu_idc_{ij}c_{ij}^{-1}
\mu_i^{-1}c_{ij} \wedge c_{ij}^{-1}\omega_ic_{ij})
\end{equation}
Also terms with no poles (i.e. terms not involving $\omega$) can
be ignored. We get
\begin{equation} \delta({\rm Tr}(-d\tilde\mu_i\tilde\mu_i^{-1}\wedge
\omega_i))(ij) = d\lambda_{ij}\lambda_{ij}^{-1}\wedge {\rm Tr}(\omega_i)=
d\lambda_{ij}\lambda_{ij}^{-1}\wedge {\rm Tr}(\omega_j)
\end{equation}
which is what we want. Note the right hand equality holds
because we are computing
mod forms regular along ${\mathcal D}$, and ${\rm Tr}(\omega_i)\equiv
{\rm Tr}(\omega_j)$ mod regular
forms by the cocycle condition \ref{coc}.
It remains to show that our cocyle \ref{coch}
does not depend on the choice
of the cocycles $\{c, \omega,\lambda, \mu\}$.
If $\lambda_{ij}$ is replaced by $\lambda_{ij}\delta(\nu)_{ij}$,
then the new \ref{coch} differs from the old one by the cochain
$$\delta\Big( \{\nu_i, {\rm det}(c_{ij})\}, d\log \nu_i \wedge
{\rm Tr}(\omega_i)\Big).$$
If $c_{ij}$ is replaced by $\gamma_ic_{ij}\gamma_j^{-1}$, then
$\omega_i$ is replaced by $d\gamma_i \gamma_i^{-1} +
\gamma_i \omega_i \gamma_i^{-1}$, and
$\mu_i$ is replaced by $\gamma_i \mu_i \gamma_i^{-1}$. The
commutativity relation implies
$d\gamma_i \mu_i \gamma_i^{-1} -\gamma_i \mu_i \gamma_i^{-1}
d\gamma_i \gamma_i^{-1}=0$. Then the new \ref{coch}
differs from the old one by
\begin{gather}
\{\{\lambda_{ij}, {\rm det}(\delta(\gamma))_{jk}\},
d\lambda_{ij} \wedge {\rm Tr} d \log (\gamma_i), \notag \\
{\rm res \ Tr} \Big(\gamma_i^{-1}d\gamma_i \omega_i
d \mu_i \mu_i^{-1} \gamma_i^{-1}
d\gamma_i -\mu_i \gamma_i^{-1} d\gamma_i \mu_i^{-1}
\gamma_i^{-1} d\gamma_i -\mu_i\gamma_i^{-1}d\gamma_i
\mu_i^{-1}\omega_i \Big) \}
\end{gather}
Using the commutativity relation \ref{comm} as well, we see that
this is this expression is the cochain
\begin{gather}
\delta\Big( \{\lambda_{ij}, {\rm det} \gamma_j\}, 0, 0\Big).
\end{gather}
\end{proof}
Now we consider the image under the map $f:C \to {\rm Spec \,}(K)$:
\begin{equation}
f_*( \{c_1(L), g, \nabla\}) \in
\Omega^1_K/d\log K^*
\end{equation}
of $\{c_1(L), g, \nabla\}$. We want to study closedness for forms in
the image of this map.
\begin{lem}\label{lem5.7} Let $\{a,b,c\}$ be a cocycle as in \eqref{coch}
representing a class in $\H^2(C,{\mathcal K}_2 \to \Omega^2_C)$,
with $db=0$. Then
$df_*\{a,b,c\}= {\rm res}_{\mathcal D}(dc)\in \Omega^2_K$.
\end{lem}
\begin{proof}
Another representative of $\{a,b,c\}$ in
$$\H^1(C, \Omega^2_C<D>({\mathcal D}') \to
\Omega^2_C<D>({\mathcal D}')/\Omega^2_C)/d\log (H^1(C, {\mathcal K}_2))$$
is of the shape $\{0, b + d\log \alpha, c\}$, thus its
derivative in
$$\H^1(C, \Omega^3_C<D>({\mathcal D}') \to
\Omega^3_C<D>({\mathcal D}')/\Omega^3_C)$$
is of the shape
$\{0, 0, dc\}$ since $db=0$. Then one applies the commutativity
of the diagramm\begin{tiny}
\begin{equation}\begin{CD}\H^1(C,\Omega^2_C\langle D\rangle({\mathcal D}') \to
\Omega^2_C\langle D\rangle({\mathcal D}')/\Omega^2_C) @> d >>
\H^1(C, \Omega^3_C\langle D\rangle({\mathcal D}') \to
\Omega^3_C\langle D\rangle({\mathcal D}')/\Omega^3_C) \\
@VVV @VVV \\
\Omega^1_K @>d>> \Omega^2_K
\end{CD}
\end{equation}
\end{tiny}
\end{proof}
\begin{thm}\label{flat}
If $\nabla^2=0$, that is if $\nabla$ is flat, then
$f_*( \{c_1(L), g, \nabla\})\in \Omega^1_{K, {\rm clsd}}/d\log
K^*$, that is the image is flat as well. In particular, this is true for
$\{c_1(\omega({\mathcal D})), \nabla\}$, as predicted by conjecture
\ref{conj}.
\end{thm}
\begin{proof}
Since, under the integrabiltiy assumption, one has in particular
$d(d\log \lambda_{ij} \wedge {\rm Tr} \omega_j)= 0$, one can apply
lemma \ref{lem5.7}.
One has to compute
\begin{gather}
\gamma= -{\rm res \ Tr}\ d(d\mu \mu^{-1} \omega) \notag \\
= -{\rm res \ Tr} (d\mu \mu^{-1})^2 \omega + {\rm res \ Tr}\ d \mu \mu^{-1}d\omega
\end{gather}
We omit the indices since we compute only with one index. Note in the
calculations which follow $\mu$ is regular and $\omega$ has poles
along ${\mathcal D}$. We write $a\equiv b$ to indicate that the polar parts of
$a$ and $b$ coincide.
The condition \ref{comm} implies
\begin{gather}
d\mu \omega + \mu d\omega\equiv d\omega \mu - \omega d\mu
\end{gather}
thus
\begin{gather}
d\mu \mu^{-1} \omega + \mu d\omega \mu^{-1}\equiv d\omega - \omega
d\mu \mu^{-1}
\end{gather}
which implies
\begin{gather}
(d\mu \mu^{-1})^2 \omega + d\mu d\omega \mu^{-1}\equiv d\mu
\mu^{-1}d\omega - d\mu \mu^{-1} \omega d\mu \mu^{-1}
\end{gather}
So taking the trace, one obtains
\begin{gather}
2 {\rm Tr} ( (d\mu \mu^{-1})^2 \omega) \equiv {\rm Tr} ((d\mu \mu^{-1} -
\mu^{-1}d\mu)d\omega)
\end{gather}
Now, under the integrability assumption
\begin{gather} \label{int}
d\omega= \omega \omega
\end{gather}
(using \eqref{comm2}) one obtains
\begin{gather}
2 {\rm Tr} ((d\mu \mu^{-1})^2 \omega) =0
\end{gather}
Now we consider the other term
\begin{gather}
\gamma'={\rm Tr} (d\mu \mu^{-1} (\omega)^2)
\end{gather}
Choosing a local basis of the bundle $E$, we write $\mu$ as a matrix
$M$, and $\omega$ as
matrix $\Omega$.
Then the condition \ref{comm}
reads
\begin{gather} \label{com2}
M \Omega= \Omega M
\end{gather}
and one has
\begin{gather}
\gamma'= {\rm Tr} (dMM^{-1}\Omega^2)
\end{gather}
The condition \ref{com2} implies
\begin{gather}
\Omega^2 = M^{-1} \Omega^2M
\end{gather}
thus
\begin{gather}
\gamma'= {\rm Tr} ( dM \Omega^2 M^{-1} )= {\rm Tr} (M^{-1}dM \Omega^2)
\end{gather}
On the other hand, differentiating the condition \ref{com2}, one
obtains
\begin{gather}
M^{-1} dM \Omega^2 + d\Omega\Omega =\notag\\
M^{-1} d\Omega M \Omega - M^{-1} \Omega dM \Omega= \notag \\
M^{-1}d\Omega \Omega M -\Omega M^{-1}dM\Omega
\end{gather}
So taking the trace, one obtains
\begin{gather}
\gamma'={\rm Tr} (M^{-1} dM \Omega^2) =
-{\rm Tr} (M^{-1}dM \Omega^2 )= -\gamma'
\end{gather}
Thus $\gamma'=0$.
\end{proof}
\section{Rank $1$ Irregular Connections in Arbitrary Dimension
on Projective Manifolds}
Let $X$ be a smooth, projective variety in characteristic 0,
and let $D\hookrightarrow X$ be a normal
crossings divisor. Given $m_i\ge 0$, define $C(X,D,\underline{m})$ to be
the group of isomorphism classes of line bundles ${\mathcal L}$ on $X$ together with
an integrable connection
$$\nabla : {\mathcal L} \to {\mathcal L}\otimes \Omega^1_X<D>({\mathcal D})
$$
where ${\mathcal D}= \sum_i m_iD_i$.
Define for $m_i\ge 1$
$$\text{Irreg}(X,D,\underline{m}) := C(X,D,\underline{m})/C(X,D,\underline{0}).
$$
\begin{thm}\label{thm6.1} With notation as above, there is a canonical
isomorphism
$$\text{Irreg}(X,D,\underline{m}) \cong \Gamma(X,{\mathcal O}_X({\mathcal D})/{\mathcal O}_X),
$$
where ${\mathcal O}_X({\mathcal D})/{\mathcal O}_X$
is the sheaf of principal parts of degree $\le m_i$ along
$D_i$.
\end{thm}
\begin{lem}Exterior differentiation induces an isomorphism
$$d: {\mathcal O}_X({\mathcal D})/{\mathcal O}_X \stackrel{\cong}{\to} \Omega^1_X<
D>({\mathcal D})_{{\rm clsd}}/
\Omega^1_X< D>_{{\rm clsd}}
$$
\end{lem}
\begin{proof}[proof of lemma] We first check injectivity. Let $x=0$ be a local
defining equation for $D$. Suppose for some $n$ with $1\le n\le m$ we had
$$d(\frac{a}{x^n}) = \frac{1}{x^n}(da - na\frac{dx}{x})
\in \Omega^1_X<D>
$$
Multiplying by $x^n$ and taking residue along $D$, it would follow that
$a|_D=0$, i.e. $\frac{a}{x^n}=\frac{b}{x^{n-1}}$.
To show surjectivity, write a local section of
$\Omega^1<D>(nD)_{{\rm clsd}}$
(here $1\le n\le m$) in the form
\begin{equation}\label{1} \omega = \frac{adx}{x^{n+1}}+\frac{B}{x^n}
\end{equation}
where $B$ does not involve $dx$. Replacing $\omega$ by
$\omega+d(\frac{a}{nx^n})$, we can assume
$$\omega = \frac{adx+B}{x^n}.
$$
Then
$$0 = d\omega = \frac{da\wedge dx + dB - n\frac{dx}{x}\wedge B}{x^n}.
$$
Multiplying by $x^n$ and taking residue along $D$, we see $B|_D = 0$. Since $B$
does not involve $dx$, it follows that $B=xC$ and $\omega$ can be written
$$\omega = \frac{adx}{x^n}+\frac{C}{x^{n-1}}
$$
Comparing with \eqref{1}, we have lowered the order of pole by $1$. This
process continues until $\omega$ has log poles.
\end{proof}
\begin{proof}[proof of theorem \ref{thm6.1}] Using the lemma, we get a diagram
with exact rows
\begin{equation*}\minCDarrowwidth.5cm \begin{CD} 0 @>>> {\mathcal O}_X^* @= {\mathcal O}_X^*
@>>> 0 @>>> 0\\
@. @VVV @VVV @VVV \\
0 @>>> \Omega^1_X<D>_{{\rm clsd}} @>>>
\Omega^1_X<D>({\mathcal D})_{{\rm clsd}} @>>>
{\mathcal O}_X({\mathcal D})/{\mathcal O}_X @>>> 0
\end{CD}
\end{equation*}
We view this as a diagram of complexes written vertically. Using the standard
hypercohomological interpretation of line bundles with connection, this yields
an exact sequence
\begin{multline*}0 \to C(X,D,0) \to C(X,D,\underline{m})
\to H^0(X,{\mathcal O}_X({\mathcal D})/{\mathcal O}_X) \\
\stackrel{\delta}{\to} \H^2(X,{\mathcal O}_X^* \to \Omega^1_X<D>_{{\rm clsd}})
\end{multline*}
We claim the map $\delta$ above is zero. In the derived category, $\delta$
factors
\begin{multline*}{\mathcal O}_X({\mathcal D})/{\mathcal O}_X \stackrel{\cong}{\to} \frac{\Omega^1_X<
D>({\mathcal D})_{{\rm clsd}}}{\Omega^1_X< D>_{{\rm clsd}}}
\stackrel{\partial}{\to} \Omega^1_X(<D>)_{{\rm clsd}}[1] \\
\to \{{\mathcal O}^*_X \to \Omega^1_X(<D>)_{{\rm clsd}} \}[2].
\end{multline*}
We have a factorization of $\partial$:
\begin{multline*}H^0({\mathcal O}_X({\mathcal D})/{\mathcal O}_X) \to H^1({\mathcal O}_X) \stackrel{\tilde d}{\to}
H^1(\Omega^1_{X, {\rm clsd}}) \to H^1(\Omega^1_X<D>_{{\rm clsd}}),
\end{multline*}
so it suffices to show the map $\tilde d$ is zero. By Hodge theory, the
composition
$$H^1({\mathcal O}_X) \stackrel{\tilde d}{\to} H^1(\Omega^1_{X, {\rm clsd}})
\stackrel{\iota}{\to} \H^1(\Omega^1_X \to \Omega^2_X)
$$
is zero, and the map $\iota$ is injective
as the complex $\{\Omega^1_X/\Omega^1_{X,{\rm clsd}} \to
\Omega^2_X\}$ is quasi-isomorphic to the complex
$\{0 \to \Omega^2_X/\Omega^2_{X, {\rm exact}}\}$, and in particular
starts in degree 1.
\end{proof}
\newpage
\bibliographystyle{plain}
\renewcommand\refname{References}
|
\section*{Introduction}
Simplicial groups and simplicial groupoids are valuable algebraic models for homotopy types.
Much has been studied about the way the group structure interacts with the simplicial structure
to yield homotopy information.
Recently the work of Wu, \cite{wut}, ~\cite{wu}, has shown that there is still progress that
can be made in calculation of homotopy invariants such as homotopy groups from simplicial groups.
Wu used techniques of combinatorial group theory, iterated commutators and properties related
to the semidirect product decompositions of the individual $G_n$ to give some insight into, for
instance, $\pi_{n+1}(\Sigma K(\pi,1))$, the homotopy groups of the suspension of an Eilenberg-MacLane
space.
Earlier Brown and Loday, \cite{bl1}, had used techniques derived from their generalised van Kampen
theorem and Loday's theory of cat\textsuperscript{$n$}-groups to give a complete description of the
$3$-type of $\Sigma K(\pi,1).$ This raises the possibility of linking the results of Wu with crossed
algebraic techniques and to combine the two techniques in order to give descriptions of, for instance,
the $k$-type of $\Sigma K(\pi,1)$ for $k=4$ and $5.$ This is still out of our
reach with the techniques of this paper, but other results
suggest the way to develop tools for this sort of task.
Carrasco \cite{carrasco}, and with Cegarra in
\cite{c:c}, gave a complete description of the extra structure of the Moore complex, ${\bf NG},$ of a
simplicial group ${\bf G}$ needed to reconstruct ${\bf G}$ from ${\bf NG}$, a sort of ultimate
generalisation of the classical Dold-Kan theorem that links simplicial abelian groups with chain
complexes. The controlled vanishing of this extra structure given necessary and sufficient conditions
for the Moore complex to be a crossed complex or crossed chain complex. Further links between simplicial
groups, their Moore complexes and crossed algebraic models for homotopy types have been given by Baues
\cite{baues1},~\cite{baues2} and \cite{baues3} and also by the second author \cite{porter}.
In this article we will develop a variant of the Carrasco - Cegarra pairing operators, that we
will call {\it Peiffer pairings}, and will show that these pairings give products of commutators, and thus,
by repeated application, iterated commutators that generate the Moore complex terms in those dimensions
where additional non-degenerate generators are not present and in general, they generate $NG_n\cap D_n$
where $D_n$ is the subgroup of $G_n$ generated by the degenerate
elements. So far it has not been possible to find a general form for the
relations between these generators. This would seem to be an extremely
hard problem in general. Some results in low dimensions and for free
simplicial groups have been obtained, but as they are incomplete they
will not be included here.
Some sample calculations of these generating elements will
be given as will some fairly elementary examples of the use of this result. \\
{\noindent\bf Acknowledgement}\\
A. Mutlu wishes to thank the University of Celal Bayar, Manisa, Republic of Turkey, for the award of a
research scholarship during the tenure of which this work was undertaken.
\section{Simplicial groups, Moore complexes and Peiffer pairings}
We refer the reader to Curtis's survey article \cite{curtis} or May's book,
\cite{may}, for most of the basic properties of
simplicial sets, simplicial groups, etc. that we will be needing.
\subsection{The Moore complex}
If ${\bf G}$ is a simplicial group, the Moore complex $({\bf NG},\partial)$ of ${\bf G}$ is the
(non-abelian) chain complex defined by
$$
NG_n = \bigcap\limits_{i=0}^{n-1}\text{Ker}d_i
$$
with $\partial_n : NG_n\longrightarrow NG_{n-1}$ induced from $d_n^n$ by restriction. It is well known
that {\em n\textsuperscript{th} homotopy group} $\pi_n({\bf G})$ of ${\bf G}$ is the
{\em n\textsuperscript{th} homology } of the Moore complex of ${\bf G}$
$$
\begin{array}{rcl}
\pi _n({\bf G}) & \cong & H_n(
{\bf NG},\partial ) \\ & = & \bigcap\limits_{i=0}^n\text{Ker}%
d_i^n/d_{n+1}^{n+1}(\bigcap\limits_{i=0}^n\text{Ker}d_i^{n+1}).
\end{array}
$$
{\bf Remark and Warning}
There is a possibility of confusion as to the exact definition of $\bf
NG$ as two conventions are currently used, one as above takes the
intersection of the $\text{Ker}d_i$ for $i < n$, the other the
intersection of the $\text{Ker}d_i$ for $0 < i \leq n$. (Curtis
\cite{curtis} uses this latter convention, whilst May, \cite{may}, uses
the former.) The two theories run parallel and are essentially `dual'
to each other, however there is a necessity for checking, which
convention is being used in any source as the actual form of any formula
usually depends on the convention being used.
\subsection{The poset of surjective maps}
We recall the following notation and terminology referring the reader
to the work of Conduch\'e, \cite{conduche}, Carrasco and Cegarra \cite{c:c} for more motivation and
some related results.
For the ordered set $[n] = \{ 0 < 1 < \cdots < n \}$, let $ {\alpha}_{i}^{n}
: [n+1] \to [n]$ be the increasing surjective map given by
\[ \alpha_{i}^{n}(j) = \left\{ \begin{array}{ll}
j & \mbox{ if $ j \leq i $ } \\
j-1 & \mbox{ if $ j > i $ }
\end{array}
\right. \]
Let $S(n , n-l)$ be the set of all monotone increasing surjective maps from
$[n]$ to $[n-l].$
This can be generated from the various $ \alpha_{i}^{n}$
by composition. The composition of these generating maps satisfies the
rule $ \alpha_{j} \alpha_{i} = \alpha_{i-1} \alpha_{j}$ with $j < i$. This
implies that every element $ \alpha \in S(n , n-l)$ has a unique expression
as $ \alpha = \alpha_{i_{1}} \alpha_{i_{2}} \ldots
\alpha_{i_{l}}$ with $0 \leq i_{1} < i_{2} < \cdots < i_{l} \leq n$, where
the indices $i_{k}$ are the elements of $[n]$ at which $\{ i_{1}, ..., i_{l},
\} = \{ i : \alpha(i) = \alpha(i+1) \}$. We thus can identify $S(n , n-l)$
with the set $\{(i_{l}, ..., i_{1}) : 0 \leq i_{1} < i_{2} < \cdots < i_{l}
\leq n-1 \}$. In particular the single element of $S(n ,n)$, defined by the
identity map on $[n]$, corresponds to the empty $0$-tuple $(~)$ denoted by
$\emptyset_{n}$. Similarly the only element of $S(n, 0)$ is $(n-1 , n-2 ,
\ldots , 0)$. For all $n \geq 0$, let
\[ S(n) = \bigcup_{0 \leq l \leq n} S(n , n-l) .\]
We say that $ \alpha = (i_{l}, ..., i_{1}) > \beta = (j_{m}, ...,
j_{1})$ in $S(n)$
\[ \mbox{ if $i_{1} = j_{1} , \cdots , i_{k} = j_{k}$ but $i_{k+1} < j_{k+1}$
$( k \geq 0 )$} \] or
\[ \mbox {if \qquad $i_{1} = j_{1} , \cdots , i_{m} = j_{m}$ and $l > m$}. \]
This makes $S(n)$ an ordered set. For instance, the orders of $S(2)$ and
$S(3)$ and $S(4)$ are respectively:
$S(2) = \{ \emptyset_{2} < (1) < (0) <(1 , 0) \}$,
$S(3) = \{ \emptyset_{3} < (2) < (1) < (2 , 1) < (0) < (2 , 0) < (1 , 0)
< (2 , 1 , 0) \}$,
$S(4) = \{ \emptyset_{4} < (3) < (2) < (3 , 2) < (1) < (3 , 1) < (2 , 1)
< (3 , 2 , 1) < (0) < (3 , 0) < (2 , 0) < (3 , 2 , 0) < (1 , 0)
< (3 , 1 , 0) < (2 , 1 , 0) < (3 , 2 , 1 , 0) \}$.
If $ \alpha, \beta \in S(n),$ we define $ \alpha \cap \beta $ to be the set of
indices which belong to both $\alpha$ and $\beta.$
If $\alpha = (i_{l}, ..., i_{1})$, then we say $\alpha$ has length $l$ and
will write $\# \alpha = l$.
\subsection{The semidirect decomposition of a simplicial group}
The fundamental idea behind this can be found in Conduch\'e \cite{conduche} .
A detailed investigation of this for the case of simplicial groups is given
in Carrasco and Cegarra \cite{c:c}.
\begin{lem}\label{1} Let ${\bf G}$ be a simplicial group. Then $G_{n}$ can be decomposed as
a semidirect product:
\[ G_{n} \cong ~\text{Ker}d_0^{n} \rtimes s_{0}^{n-1}(G_{n-1}) \]
\end{lem}
\begin{pf} The isomorphism can be defined as follows:
\[ \theta : G_{n} \to \text{Ker}d_{0}^{n}\rtimes s_{0}^{n-1}(G_{n-1}) \]
\[ g \mapsto (g s_{0}d_{0}g^{-1}, s_{0}d_{0}g) .\]
\hfill\end{pf}\\
Since we have the isomorphism $G_{n} \cong ~\text{Ker}d_{0} \rtimes s_{0}G_{n-1}$,
we can repeat this process as often as necessary to get each of the $G_{n}$
as a multiple semidirect product of degeneracies of terms in the Moore
complex. In fact, let ${\bf K}$ be the simplicial group defined
by
$$
\begin{array}{cccc}
K_n=\text{Ker}d_{0}^{n+1}, & d_i^n=d_{i+1}^{n+1}\mid _{\text{Ker}
d_{0}^{n+1}} & \text{and} & s_i^n=s_{i+1}^{n+1}\mid _{\text{Ker}
d_{0}^{n+1}}.
\end{array}
$$
Applying Lemma ~\ref{1} above, to $G_{n-1}$ and to $K_{n-1}$, gives
$$
\begin{array}{rcl}
G_n & \cong & \text{Ker}d_0\rtimes s_{0}G_{n-1} \\
& = & \text{Ker}d_0 \rtimes s_{0}(\text{Ker}d_{0}\rtimes s_{0}G_{n-2}) \\
& = & K_{n-1}
\rtimes \ (s_{0}\text{Ker}d_{0}\rtimes s_{0}s_{0}G_{n-2}).
\end{array}
$$
Since ${\bf K}$ is a simplicial group, we have the following
$$
\begin{array}{rcl}
\text{Ker}d_0=K_{n-1} & \cong & \text{Ker}d_{0}^K\rtimes s_{0}^KK_{n-2} \\
& = & (\text{Ker}d_{1}\cap \text{Ker}d_0)\rtimes s_{1}\text{Ker}d_{0}
\end{array}
$$
and this enables us to write
$$
G_n=((\text{Ker}d_{1}^n\cap \text{Ker}d_0^n)\rtimes
s_{1}(\text{Ker}d_{0}^{n-1}))\rtimes
(s_{0}(\text{Ker}d_{0}^{n-1})\rtimes s_{0}s_{0}(G_{n-2})).
$$
We can thus decompose $G_{n}$ as follows:
\begin{prop}\label{2} (cf. \cite{conduche}, p.158) If ${\bf G}$ is a simplicial group, then for any $n \geq 0$
$$
\begin{array}{lll}
G_n & \cong & (\ldots (NG_n
\rtimes s_{n-1}NG_{n-1})\rtimes \ldots \rtimes s_{n-2}\ldots s_1NG_1)\rtimes
\\& &\quad(\ldots (s_{0}NG_{n-1}\rtimes s_{1}s_{0}NG_{n-2})%
\rtimes \ldots \rtimes s_{n-1}s_{n-2}\dots s_0NG_0).~\square
\end{array}
$$
\end{prop}
The bracketing and the order of terms in this multiple semidirect product
are generated by the sequence:
$$
\begin{array}{lll}
G_1 & \cong & NG_1
\rtimes s_0NG_0 \\ G_2 & \cong & (NG_2
\rtimes s_1NG_1)\rtimes (s_0NG_1\rtimes s_1s_0NG_0) \\ G_3 & \cong & ((NG_3
\rtimes s_2NG_2)\rtimes (s_1NG_2\rtimes s_2s_1NG_1))\rtimes \\& &\qquad
((s_0NG_2\rtimes s_2s_0NG_1)\rtimes (s_1s_0NG_1\rtimes %
s_2s_1s_0NG_0)).
\end{array}
$$
and
$$
\begin{array}{lll}
G_4 & \cong & (((NG_4
\rtimes s_3NG_3)\rtimes (s_2NG_3\rtimes s_3s_2NG_2))\rtimes \\ & & \qquad
\ ((s_1NG_3
\rtimes s_3s_1NG_2) \rtimes(s_2s_1NG_2\rtimes s_3s_2s_1NG_1)))\rtimes \\ &
& \qquad \qquad s_0( \text{decomposition of }G_3).
\end{array}
$$
Note that the term corresponding to $\alpha
=(i_l,\ldots ,i_1)\in S(n)$ \ \ is \ \
$s_\alpha (NG_{n-\#\alpha
})=s_{i_l...i_1}(NG_{n-\#\alpha })=s_{i_l}...s_{i_1}(NG_{n-\#\alpha }),$
where $\#\alpha =l.$
Hence any element $x\in G_n$ can be written in the form%
$$
x=y \prod\limits_{\alpha \in S(n)}s_\alpha (x_\alpha )\text{ \qquad with }%
y\in NG_n\text{ and }x_\alpha \in NG_{n-\#\alpha }.
$$
\section{Peiffer pairings generate}
In the following we will define a normal subgroup $N_n$ of $G_n.$ First of all we
adapt ideas from Carrasco \cite{carrasco} to get the construction of a useful
family of natural pairings.
We define a set $P(n)$ consisting of pairs of elements $(\alpha , \beta)$
from $S(n)$ with $\alpha \cap \beta =\emptyset $ and $\beta < \alpha $, with respect to
lexicographic ordering in $S(n)$ where $\alpha =(i_l,\ldots ,i_1),\beta=(j_m,...,j_1)\in S(n).$
The pairings that we will need,
$$
\{F_{\alpha ,\beta }:NG_{n-\#\alpha}\times NG_{n-\#\beta}\longrightarrow
NG_n:(\alpha ,\beta )\in P(n),\ \ n\geq 0\}
$$
are given as composites by the diagram
$$
\diagram
NG_{n-\#\alpha} \times NG_{n-\#\beta} \rto^{\hspace{.9cm} F_{\alpha, \beta}}
\dto_{s_{\alpha} \times s_{\beta}} & NG_n \\
G_n \times G_n \rto^{\mu} & G_n \uto_p
\enddiagram
$$
where
$$
\begin{array}{c}
s_\alpha =s_{i_l}\ldots s_{i_1}:NG_{n-\#\alpha}\longrightarrow G_n, \ \
s_\beta =s_{j_m}\ldots s_{j_1}:NG_{n-\#\beta}\longrightarrow G_n,
\end{array}
$$
$p:G_n\rightarrow NG_n$ is defined by the composite projections $p(x)=p_{n-1}\ldots
p_0(x),$ where
$$
p_j(z) = zs_jd_j(z)^{-1}~\qquad \text{with\quad }j=0,1,\ldots, n-1,
$$
$\mu :G_n\times G_n\rightarrow G_n$ is given by the commutator map and
$\#\alpha$ is the number of the elements in the set of $\alpha$, similarly for
$\#\beta.$ Thus
$$
\begin{array}{rcl}
F_{\alpha ,\beta }(x_\alpha , y_\beta ) & = & p\mu (s_\alpha \times
s_\beta )(x_\alpha , y_\beta ) \\
& = & p[s_\alpha x_\alpha , s_\beta y_\beta ] .
\end{array}
$$
\textbf{Definition} Let
$N_n$ or more exactly $N_n^G$ be the normal subgroup of $G_n$ generated by elements of
the form
$$
F_{\alpha ,\beta }(x_\alpha , y_\beta )
$$
where $x_\alpha \in NG_{n-\#\alpha}$ ~and ~ $y_\beta \in NG_{n-\#\beta}$.
\medskip
This normal subgroup $N_n^G$ depends functorially on $G$, but we will usually
abbreviate $N_n^G$ to $N_n$, when no change of group is involved.
\medskip
We illustrate this subgroup for $n=2$ and $n=3$ to show what it looks
like.\\
\textbf{Example (a) :}
For $n=2,$ suppose $\alpha =(1)$, $\beta =(0)$ and \\
$x_1,y_1\in NG_1={\rm Ker}d_0$. It follows that
$$
\begin{array}{rcl}
F_{(0)(1)}(x_1 , y_1) & = & p_1p_0[s_0x_1 , s_1y_1] \\
& = & p_{1}[s_0x_1, s_1y_1] \\
& = & [s_0x_1, s_1y_1]{~} [s_1y_1, s_1x_1]
\end{array}
$$
is a generating element of the normal subgroup $N_2.$
For $n=3,$ the possible pairings are the following \label{sayfa}
$$
\begin{array}{lll}
F_{(1,0)(2)}, & F_{(2,0)(1)}, & F_{(0)(2,1)}, \\
F_{(0)(2)}, & F_{(1)(2)}, & F_{(0)(1)}.
\end{array}
$$
For all $x_1\in NG_1,~y_2\in NG_2,$ the corresponding generators of $N_3$ are:\label{sayfa1}
$$
\begin{array}{lll}
F_{(1,0)(2)}(x_1, y_2) & = & [s_1s_0x_1, s_2y_2]{~} [s_2y_2, s_2s_{0}x_1] \\
F_{(2,0)(1)}(x_1, y_2) & = & [s_2s_0x_1, s_1y_2]{~} [s_1y_2, s_2s_1x_1]{~}
[s_2s_1x_1, s_2y_2]{~} [s_2y_2, s_2s_0x_1]
\end{array}
$$
and all $x_2\in NG_2,~y_1\in NG_1,$
$$
\begin{array}{lll}
F_{(0)(2,1)}(x_2, y_1) & = & [s_0x_2, s_2s_1y_1]{~}[s_2s_1y_1, s_1x_2]{~}
[s_2x_2, s_2s_1y_1]
\end{array}
$$
whilst for all $x_2 , y_2 \in NG_2$,\label{sayfa1b}
$$
\begin{array}{lll}
F_{(0)(1)}(x_2 , y_2) & = & [s_0x_2 , s_1y_2]{~} [s_1y_2 , s_1x_2]{~} [s_2x_2 ,s_2y_2] \\
F_{(0)(2)}(x_2 , y_2) & = & [s_0x_2 , s_2y_2] \\
F_{(1)(2)}(x_2 , y_2) & = & [s_1x_2 , s_2y_2]{~}[s_2y_2 , s_2x_2].
\end{array}
$$
Our aim in this paper is to prove that the images of these pairings generate
$NG_n\cap D_n.$ More precisely:
\begin{thm}~~{\bf ( Theorem A )}\label{10}
Let {\bf G} be a simplicial group and for $n>1,$ let $D_n$ the
subgroup in $G_n$ generated by degenerate elements.
Let $N_n^G$ be the normal subgroup generated by elements of the form
$$
F_{\alpha ,\beta }(x_\alpha , y_\beta )\qquad \text{with }(\alpha
,\beta )\in P(n)
$$
where $x_{\alpha} \in NG_{n-\#\alpha },\ y_{\beta} \in NG_{n-\#\beta }$.
Then
$$
NG_n\cap D_n = N_n^G\cap D_n.
$$
\end{thm}
As a corollary we, of course, have that the image of $N_n^G\cap D_n$ is equal to the image of $NG_n\cap D_n$
i.e., $\partial _n(N_n\cap D_n)=\partial _n(NG_n\cap D_n)$.\\
The proof of 2.1 is given in the next section after some preparatory lemmas. Here we
restrict to the case $n=2$ by way of illustration. In their paper \cite{bl1},
Brown and Loday proved a lemma:
\begin{lem}~\cite{bl1}
Let ${\bf G}$ be a simplicial group such that $G_2 = D_2$ is generated by
degenerate elements. Then in the Moore complex of ~${\bf NG}$ we have
$\partial_2NG_2 = \partial_2N_2$ where $N_2$ is the normal subgroup of $G_2$ generated
by elements of the form
$$
\begin{array}{lll}
F_{(0)(1)}(x_1,y_1) &=& [s_0x_1,s_1y_1]{~}[s_1y_1, s_1x_1]\\
\end{array}
$$
with $x_1,y_1\in NG_1.$
\end{lem}
This is, of course, a trivial consequence of Theorem A and their proof
inspired that of the more general theorem given here.
\textbf{Remark:}
AN unknown referee made the interesting observation that if $x$ is a Moore
cycle, so $\partial x = 0$, then $F_{(0)(1)}(x,x)$ is one also. Thus
$F_{(0)(1)}$ induces an operation $\pi_\ast(G) \rightarrow \pi_{\ast +1}(G)$.
Geometrically this operation can be described as the $\eta$-operation given by
the composition $S^{m+1}\stackrel{\eta}'{\rightarrow}S^m
\stackrel{x}{\rightarrow}G$, where $\eta$ is the (suspension of) the Hopf map.
The geometric interpretation of the $F_{\alpha \beta(1)}$ in general would
seem to be quite important but the authors have as yet little idea what it
might be.
\section{Elements of $N_n$ and properties of the pairings}
In the following we analyse various types of elements in $N_n$ and show that
products of them give elements that we want in giving an alternative description of
$NG_n$.
\begin{lem}\label{7}
Given $x_\alpha \in NG_{n-\#\alpha },~y_\beta \in NG_{n-\#\beta }$ with $%
\alpha =(i_l,\ldots ,i_1),$ $\beta =(j_m,\ldots ,j_1)\in S(n).$ If $\alpha
\cap \beta =\emptyset $ with $\beta <\alpha $ and $v=[s_\alpha x_\alpha,
s_\beta y_\beta],$ then
(i) \ \ if $k\leq i_1,$ then $p_k(v)=v,$
(ii)\ \ \thinspace if $k>i_l+1\,$or $k>j_m+1,$ then $p_k(v)=v,$
(iii)\ \ if $k\in \{j_1,\ldots ,j_m \}$ and $k=i_r+1$ for some $r,$ then for \\
${\alpha'} = (i_l,\ldots,i_r+1,i_r-1,\ldots, i_1)$ and
${\beta} = (j_m,j_{m-1},\ldots, j_1),$
$$
p_k(v)=[s_\alpha x_\alpha, s_\beta y_\beta]{~}[s_{\alpha^{\prime}}
x_{\alpha}, s_{\beta}y_{\beta}]^{-1},
$$
(iv)\ \ \ if $k\in \{i_1, \ldots, i_l \}$ and $k=j_s+1$ for some $s,$ then
for \\
${\beta'} = (j_m,\ldots,j_s+1,j_s-1\ldots, j_1)$
$$
\begin{array}{lll}
p_k(v) & = & [s_\alpha x_\alpha, s_\beta y_\beta ]{~}
[s_{\alpha}x_{\alpha}, s_{\beta^{\prime}}y_{\beta}]^{-1} \\
& = & vv^{\prime}
\end{array}
$$
where $v^{'}\in G_{n-1}$ and $0\leq k\leq n-1,$
(v)\ \ \ if $k = j_m+1$ (or $k = i_l+1$) then
$$
\begin{array}{lll}
p_k(v) & = & vs_k(v_k)^{-1}\\
& = & [s_\alpha x_\alpha, s_\beta y_\beta]s_k(v_k)^{-1}
\end{array}
$$
where $s_k(v_k)^{-1} = [s_{\beta^{\prime}}y_{\beta}, s_{\alpha}x_{\alpha}]$
$(\text{or}\quad s_k(v_k)^{-1} = [s_{\beta}y_{\beta}, s_{\alpha^{\prime}}x_{\alpha}])$
with respect to $k= j_m+1$ (and $k= i_l+1$ respectively) and for new strings
${\alpha'}$ and ${\beta'},$
(vi)\ \ \ if $k = j_1+1,$ then
$$
\begin{array}{lll}
p_k(v) & = & vs_k(v_k)\\
& = & [s_\alpha x_\alpha, s_\beta y_\beta]s_k(v_k)^{-1}
\end{array}
$$
where $s_k(v_k)^{-1}=[s_{\beta^{\prime}}y_{\beta}, s_{\alpha}x_{\alpha}]$
(vii)\ \ \ if $k \in \{j_1,\ldots,j_m,j_{m+1}\}$ and $k=i_t+1$ for some $t,$ then
$$
\begin{array}{lll}
p_k(v) & = & [s_\alpha x_\alpha, s_\beta y_\beta]
[s_{\beta'}y_\beta, s_{\alpha'}x_\alpha]
\end{array}
$$
where $0\leq k \leq n-1.$
\end{lem}
\begin{pf}
Assume $\beta <\alpha $ and $\alpha \cap \beta =\emptyset $ which implies $%
i_1<j_1.$ In the range $0\leq k\leq i_1,$
$$
\begin{array}{lll}
p_k(v) & = & [s_\alpha x_\alpha, s_\beta y_\beta ]{~}[s_kd_ks_{i_{m}}\ldots
s_{i_{1}}x, s_kd_ks_{j_{m}}\ldots s_{j_{1}}y]^{-1} \\
& = & [s_\alpha (x_\alpha ), s_\beta (y_\beta )]{~}[s_{j_{m}-1}\ldots s_{j_{1}-1}s_kd_ky,
s_{i_{m}-1} \ldots s_{i_{1}-1}s_kd_kx] \\
& = & [s_\alpha x_\alpha, s_\beta y_\beta]\qquad \text{since }d_k(x_\alpha
)=1 \ \ or \ \ d_k(y_\beta)=1. \\
& = & v .\\
\end{array}
$$
Similarly if $k>i_l+1,$ then%
$$
\begin{array}{lll}
p_k(v) & = & [s_\alpha x_\alpha, s_\beta y_\beta ]{~}[s_{j_{m}}\ldots s_{j_{1}}
s_{k-m}d_{k-m}y, s_{i_{m}}\ldots s_{i_{1}}s_{k-m}d_{k-m}x] \\
& = & [s_\alpha x_\alpha, s_\beta y_\beta ]\qquad \text{since }%
d_{k-m}(y_\beta )=1 \ \ or \ \ d_{k-l}(x_\alpha)=1.\\
& = & v .
\end{array}
$$
Clearly the same sort of argument works if $k>j_{m+1}.$
\noindent If $k\in \{j_1,\ldots ,j_m,j_{m+1} \}$ and $k=j_t+1$ for some $t,$
then%
$$
\begin{array}{lll}
p_k(v) & = & [s_\alpha x_\alpha, s_\beta y_\beta ]{~}[s_kd_ks_\beta
y_\beta , s_kd_ks_\alpha x_\alpha] \\
& = & [s_\alpha x_\alpha, s_\beta y_\beta]{~}[s_{\beta^{\prime}}
x_{\beta^{\prime}}, s_{\alpha^{\prime}}y_{\alpha^{\prime}}]\\
\end{array}
$$
\hfill\end{pf}
\begin{lem}\label{8}
If $\alpha \cap \beta =\emptyset $ and $ \beta<\alpha ,$ then
$$
p_{l}\ldots p_1{~} [s_\alpha x_\alpha , s_\beta y_\beta]= [s_\alpha x_\alpha
, s_\beta y_\beta ] \prod\limits_{i=1}^{l}s_i(z_i)^{-1}
$$
where $z_i \in \bigcap\limits_{j=0}^{i-1}$\text{\rm Ker}$d_j\subset G_{n-1}$ and $l\in [n-1].$
\end{lem}
\begin{pf}
By induction on $l.$
\end{pf}
\begin{lem}\label{9}
Let $x_\alpha \in NG_{n-\#\alpha },~y_\beta \in NG_{n-\#\beta }$ with $%
\alpha ,\beta \in S(n),$ then
$$
s_\alpha x_\alpha s_\beta y_\beta s_\alpha (x_\alpha )^{-1}
=s_{\alpha \cap \beta }z_{\alpha \cap
\beta }
$$
where $z_{\alpha \cap \beta }$ has the form $s_{\bar{\alpha}}x_\alpha
s_{\bar{\beta}}y_\beta s_{\bar{\alpha}}(x_\alpha)^{-1}
$ and $\bar{\alpha} \cap \bar{\beta}=\emptyset .$
\end{lem}
\begin{pf}
If $\alpha \cap \beta =\emptyset ,$ then this is trivially true. Assume $%
\#(\alpha \cap \beta )=t,$ with $t\in {\mathbb N}.$ Take $\alpha =(i_l,\ldots ,i_1)$ ~and ~$%
\beta =(j_m,\ldots ,j_1)$ with $\alpha \cap \beta =(k_t,\ldots ,k_1),$
$$
s_\alpha x_\alpha = s_{i_l}\ldots s_{k_t}\ldots s_{i_1}x_\alpha ~~~\text{%
and}~~~s_\beta y_\beta = s_{j_m}\ldots s_{k_t}\ldots s_{j_1}y_\beta .
$$
Using repeatedly the simplicial axiom $s_as_b=s_bs_{a-1}$ for $b<a$ until
$s_{k_t}\ldots s_{k_1}$ is at the beginning of the string, one
gets the following
$$
s_\alpha x_\alpha = s_{k_t\ldots k_1}(s_{\bar{\alpha} }x_\alpha )~~~
\text{and}~~~s_\beta y_\beta = s_{k_t\ldots k_1}(s_{\bar{\beta}}y_\beta ).
$$
Multiplying these expressions together gives
$$
\begin{array}{lll}
s_\alpha x_\alpha s_\beta y_\beta s_\alpha (x_\alpha )^{-1}
& = & s_{k_t}\ldots s_{k_1}(s_{\bar{\alpha}}
x_\alpha )s_{k_t}\ldots s_{k_1}(s_{\bar{\beta}}y_\beta )
s_{k_t}\ldots s_{k_1}(s_{\bar{\alpha}}(x_\alpha)^{-1}) \\
& = & s_{k_t\ldots k_1}(s_{\bar{\alpha}}x_\alpha s_{\bar{\beta}}
y_\beta s_{\bar{\alpha}}(x_\alpha)^{-1})\\
& = & s_{\alpha \cap \beta }(z_{\alpha \cap \beta }),
\end{array}
$$
where $z_{\alpha \cap \beta }= s_{\bar{\alpha}}x_\alpha s_{\bar{\beta}}
y_\beta s_{\bar{\alpha}}(x_\alpha)^{-1}
\in {NG_{n-\#(\alpha \cap \beta )}}$ and where
$\bar{\alpha} = (i_l-t,\ldots, k_t+1-t,\dots,i_1)$ and
$\bar{\beta} =(j_m-t,\ldots, {k_t'}+1-t,\dots,j_1).$
Hence $\bar{\alpha} \cap \bar{\beta}
=\emptyset_{n-\#(\alpha \cap \beta )}.$
Moreover $\bar{\alpha} <\alpha $ and $~\bar{\beta}
<\beta $ as $\#\bar{\alpha}<\#\alpha $ and $~\#\bar{\beta}
<\#\beta.$
\end{pf}
Suppose $\alpha= (i_s,\ldots,i_1)\in S(m)$ and $\gamma:[n]\longrightarrow
[m]\in S(n,m)$. Define $\gamma_{\ast}(\alpha)$ by
$s_{\gamma_{\ast}(\alpha)}= s_{\gamma}s_{\alpha}$.
\begin{cor}\label{baba}
Let $\beta \leq \alpha$ and
$\gamma:[n]\longrightarrow [m].$
Then $\gamma_{{\ast}(\beta)}\leq\gamma_{{\ast}(\alpha)}\Longleftrightarrow\beta\leq\alpha,$
where $\gamma_{{\ast}(\alpha)},\ \gamma_{{\ast}(\beta)}\in S(n).$ $\square$
\end{cor}
The following lemma is proved similarly.
\begin{lem}\label{anne}
For $m\leq n,$ suppose given in $G_m$ an element
$$
g = \prod\limits_{{\beta'}\leq{\gamma'}\leq{\alpha'}}s_{\gamma'}(z_{\gamma'})
$$
and $s_{\delta}: G_m\longrightarrow G_n.$ Then setting $\alpha, \ \beta\in S(n)$
such that\\
$s_{\delta}s_{\alpha'} = s_{\alpha}, \ s_{\delta}s_{\beta'} = s_{\beta}$
$$
s_{\delta}(g) = \prod\limits_{{\beta}\leq{\gamma}\leq{\alpha}}s_{\gamma}(z_{\gamma})
$$
for some elements $z_{\gamma}\in NG_{n-\#{\gamma}}$ and where
$s_{\delta}s_{\gamma'} = s_{\gamma}.$ $\square$
\end{lem}
{\noindent \bf Proof of Theorem A :}\\
From Proposition \ref{2}, $G_n$ is isomorphic to%
$$
NG_n\rtimes s_{n-1}NG_{n-1}\rtimes s_{n-2}NG_{n-1}\rtimes \ldots \rtimes
s_{n-1}s_{n-2}\dots s_0NG_0.
$$
Similarly
$D_n$ is isomorphic to\\
$(NG_n\cap D_n)\rtimes s_{n-1}NG_{n-1}\rtimes s_{n-2}NG_{n-1}\rtimes \ldots \rtimes
s_{n-1}s_{n-2}\dots s_0NG_0.$~~
Hence any element $g$ in $D_n$ can be written in the following form%
$$
g = g_ns_{n-1}(y_{n-1})s_{n-2}(y_{n-1}^{\prime
})s_{n-1}s_{n-2}(y_{n-2})\ldots s_{n-1}s_{n-2}\ldots s_0(y_0),\text{\quad}
$$
with $g_n\in NG_n\cap D_n,$ $y_{n-1},y_{n-1}^{\prime }$ $\in NG_{n-1},$ $y_{n-2}\in
NG_{n-2},$ $y_0$ $\in NG_0$ etc.\\
To simplify the notation a little, we will assume that $G_n = D_n$, so that
$N_n \subset D_n$. The general case would replace $NG_n$ by $NG_n\cap D_n$ and
similarly $N_n$ by $N_n\cap D_n$ from here on.
As it is easily checked that $N_n\subseteq NG_n\cap D_n,$
it is enough to prove that any element in $D_n/N_n$ can be
written in the form
$$
s_{n-1}(y_{n-1})s_{n-2}(y_{n-1}^{\prime })s_{n-1}s_{n-2}(y_{n-2})\ldots
s_{n-1}s_{n-2}\ldots s_0(y_0)N_n
$$
that is, for any $g\in D_n$,
$$
gN_n=s_{n-1}(y_{n-1})s_{n-2}(y_{n-1}^{\prime })\ldots
s_{n-1}s_{n-2}\ldots s_0(y_0)N_n.
$$
for some $y_{n-1}\in NG_{n-1},$ etc. We refer to this as the standard form of $gN_n.$ \newline
If $g\in D_n,$ it is a product of degeneracies.
If $g$ is itself a degenerate element, it is obvious
that it is a product of elements in the semidirect factors,
$s_\beta (G_{n-\#\beta }),$ ~$\beta\in S(n)- \{\emptyset_n\}.$
\newline
Assume therefore that
provided an element $g$ can be written as a product of $k-1$ degeneracies of
this form, then it has the desired form modulo $N_n.$ Now for an element $g$ which
needs $k$ degenerate elements, we have
$$
g=s_{\alpha}x_\alpha g^{\prime }
\qquad \text{with }x_\alpha \in NG_{n-\#\alpha}
$$
where $g^{\prime }$ needs fewer than $k$ and so%
$$
\begin{array}{lll}
gN_n & = & s_\alpha x_\alpha g^{\prime }N_n \\
& = & s_\alpha x_\alpha (s_{n-1}(y_{n-1})s_{n-2}(y_{n-1}^{\prime })\ldots
s_{n-1}s_{n-2}\ldots s_0(y_0))N_n.
\end{array}
$$
We prove that this can be rewritten in the desired form $\text{mod}~N_n$
by using induction on $\alpha$ within the linearly ordered set
$S(n)- \{\emptyset_n\}.$ \\
If $\alpha =(n-1)$, then
$$
gN_n = s_{n-1}(xy_{n-1})s_{n-2}(y_{n-1}^{\prime })\ldots
s_{n-1}s_{n-2}\ldots s_0(y_0) N_n
$$
where $x\in NG_{n-1}$ and $(xy_{n-1})\in NG_{n-1}.$ \\
If $\alpha =(n-2),$ then since
$$
F_{(n-2)(n-1)}(x_{n-1}, y_{n-1}) = [s_{n-2}x_{n-1},~ s_{n-1}y_{n-1}]
~[s_{n-1}y_{n-1},~ s_{n-1}x_{n-1}]
$$
and
$$
s_{n-2}(x_{n-1})s_{n-1}(y_{n-1})s_{n-2}(x_{n-1})^{-1} \equiv s_{n-1}(x_{n-1}y_{n-1}x^{-1}_{n-1})
\quad\text{mod}~N_n,
$$
we have
$$
\begin{array}{lll}
gN_n & =& (s_{n-2}(x_{n-1})s_{n-1}(y_{n-1})s_{n-2}(x_{n-1})^{-1})
s_{n-2}(x_{n-1})s_{n-2}(y_{n-1}^{\prime }) \\
& & \ldots s_{n-1}s_{n-2}\ldots s_0(y_0) N_n \\
& = & s_{n-1}(xyx^{-1}_{n-1})s_{n-2}(xy_{n-1}^{\prime })\ldots
s_{n-1}s_{n-2}\ldots s_0(y_0) N_n
\end{array}
$$
where $x_{n-1}, y_{n-1}\in NG_{n-1}$ so $(xyx_{n-1}^{-1}), ~~
(xy_{n-1}^{\prime })\in NG_{n-1}.$ \\
In general we need to sort $s_{\alpha}x_{\alpha}$ into its correct
place in the product but in so doing will conjugate earlier terms in the product
as happened in the case $\alpha= (n-2)$ above. Each of these terms must be
shown to consist only of subterms of types we have already dealt with, that is
further to the left in the standard form of the product. Explicitly we assume
that we can do this sorting for any term $s_{\gamma}x_{\gamma}$ with
$\gamma < \alpha$ and examine
$$
\begin{array}{lll}
gN_n & =& s_{\alpha}x_{\alpha}(s_{n-1}(y_{n-1})s_{n-2}(y_{n-1}^{\prime })\ldots
s_{n-1}s_{n-2}\ldots s_0(y_0))N_n \\
& = &s_{\alpha}x_{\alpha}
\prod\limits_{\beta\in S(n)-\{\emptyset_{n}\}}s_{\beta}y_{\beta}N_n \\
gN_n & = &\prod\limits_{\alpha>\beta}s_\alpha x_\alpha s_\beta y_{\beta}s_\alpha (x_\alpha )^{-1}
\cdot s_\beta (xy)_\beta \cdot~\prod\limits_{\beta >\alpha}s_\beta y_{\beta}N_n
\end{array}
$$
where $\beta\in S(n)-{\emptyset_n}$ and $\alpha>\beta$ with respect to
the lexicographic ordering in $S(n).$ \\
We look at products of the following type
$$
s_\alpha x_\alpha s_\beta y_\beta
s_\alpha (x_\alpha)^{-1}\qquad (\ast)
$$
and we want to show that these can always be written in the form
$$
\prod\limits_{\gamma\leq\beta}s_{\gamma}(z_{\gamma})
$$
for some $z_{\gamma}\in NG_{n-\#\gamma}.$ This will mean that
we already know how to sort all the terms that arise since none occur
`to the right of' $\beta$ in the lexicographic order in the product.\\
We check this product case by case as follows: \\
If $\alpha\cap\beta = \emptyset,$ then by Lemma~\ref{8},
$$
s_\alpha x_\alpha s_\beta y_\beta s_\alpha (x_\alpha )^{-1}
\equiv\prod\limits_{k=l}^{i_1+1}s_k(z_k)s_\beta y_\beta \quad\text{mod}~N_n,
$$
where $\beta\in S(n)-\{\emptyset_n\}.$
Now we need to show that each $s_k(z_k)$ is made up of terms $s_\mu(z_\mu)$
with $(z_\mu)\in NG_{n-\#\mu},$ $\mu\leq\beta.$ (We will use the notation of Lemma ~\ref{7}.)
Since $\alpha>\beta$ then $i_1\leq j_1.$ We have
$$
z_k = \prod\limits_{\mu\leq(k-1)}s_\mu(z_\mu)
\quad\text{since}\quad z_k\in \bigcap\limits_{j=0}^{k-1}\text{Ker}d_j
$$
so
$$
s_k(z_k) = \prod\limits_{\mu\leq(k-1)\leq(i_1)}s_ks_\mu(z_\mu)
$$
where we write $\mu =(m_1,\ldots,m_r)$ so we have $k-1\leq m_1$ or $k\leq m_1+1.$ \\
We compare $k$ with $m_1$ and $m_2$: either \\
(a) ~$k = m_1+1 < m_2;$ \\
(b) ~$k = m_1+1 = m_2 ;$ \\
$\spreaddiagramrows{-1.2pc} \spreaddiagramcolumns{-1.2pc}
\def\ssize} \def\labelstyle{\ssize{\ssize} \def\labelstyle{\ssize}
(c) ~k=m_1 = \left\{\begin{array}{ll}
\text{and}& m_2 =m_1+1, \text{or}\\
\text{and}& m_2 > m_1+1,
\end{array}\right.$\\
or \\
(d) $k< m_1.$ \\
Thus \\
\begin{equation*}
s_ks_{\mu}= s_{m_r+1}\ldots s_ks_{m_2}s_{m_1} =\left\{\begin{array}{lll}
s_{m_r+1}\ldots s_{m_2+1}s_ks_{m_1}& \quad\text{cases (a) and (b),}\\
s_{m_r+1}\ldots s_{m_1+1}s_{m_1} &\quad\text{case (c),}\\
s_{m_r+1}\ldots s_{m_1+1}s_{k} &\quad\text{case (d),}
\end{array}\right.
\end{equation*}
so in each case $s_ks_{\mu} = s_{\mu'}$ where ${m_1'} = \text{min}\{k,m_1\}.$
We compare ${\mu'}$ with $\beta.$ If ${m_1'} > j_1,$ then ${\mu'}\leq \beta.$
If ${m_1'} =j_1,$ then either $k=j_1$ or $m_1 =j_1$ then $k = j_1+1.$ Thus we need
to show
$$
w=s_{j+1}d_{j+1}[s_{\beta} y_{\beta} ,~s_{\alpha} x_{\alpha} ] =
\prod\limits_{\vartheta\leq\beta}s_{\vartheta}(z_{\vartheta}).
$$
There are two cases:\\
(i){~}If $j_1\in \alpha,$ then
$$
w=[s_{\beta'} y_{\beta} ,~s_{\alpha} x_{\alpha} ] =
\prod\limits_{\vartheta\leq{\beta'}\leq\beta}s_{\vartheta}(z_{\vartheta})
$$
where $\beta\leq{\beta'}$ and since $j_1+1 =j_s\notin\beta,$ then
${\beta'} = \{j_{m},\ldots, j_{s+1}+1, j_s, j_{s-1}, \ldots,j_1\},$ \\
(ii){~}If $j_1+1\in \beta,$ then
$$
w=[s_{\beta} y_{\beta} ,~s_{\alpha'} x_{\alpha} ] =
\prod\limits_{\vartheta\leq\beta}s_{\vartheta}(z_{\vartheta})
$$
where $\alpha\leq{\alpha'}$and since $j_1+1= i_r\notin\alpha,$ then
${\alpha'} = \{i_{l}, \ldots, i_{r+1}+1, i_r, i_{r-1}, \ldots,i_1\}.$ \\
Both cases are covered by the induction hypothesis. Both cases can thus be written
$$
\prod\limits_{\vartheta\leq\beta}s_{\vartheta}(z_{\vartheta}).
$$
We have $z_{\vartheta}\in NG_{n-\#\vartheta}$ and for
some $r$ and $s$ then it can be written for above cases
$$
s_k(z_k) = \prod\limits_{{\emptyset}_n\leq{\gamma'}\leq(k)\leq(i_1)}s_{\gamma'}(z_{\gamma})
$$
where $s_ks_\gamma = s_{\gamma'},$ ${\gamma'}\leq\beta$ and $z_{\gamma'}\in NG_{n-\#\gamma'}$
so
$$
gN_n = \prod\limits_{{\gamma'}\leq\beta}s_{\gamma'}(z_{\gamma'})\cdot s_{\alpha}(xy_{\alpha})\cdot
\prod\limits_{{\alpha}<\beta}s_{\beta} y_{\beta}N_n
$$
as required.\\
If $\alpha \cap \beta \neq \emptyset $, then one gets, from Lemma \ref{9},
the following
$$
s_\alpha x_\alpha s_\beta y_\beta s_\alpha (x_\alpha)^{-1}
= s_{\alpha \cap \beta }(s_{\bar{\alpha}} x_\alpha s_{\bar{\beta}} y_\beta
s_{\bar{\alpha}}(x_\alpha )^{-1})
$$
where $\bar{\alpha}> \bar{\beta},$ $\bar{\alpha}\cap\bar{\beta}\in \emptyset_{n-\#(\alpha \cap \beta)}.$
Using Lemma ~\ref{anne} in dimension $n-\#(\alpha \cap \beta)$ and Corollary ~\ref{baba} then we have
$$
(s_{\bar{\alpha}} x_\alpha s_{\bar{\beta}} y_\beta s_{\bar{\alpha}}(x_\alpha )^{-1})\equiv
\prod\limits_{\theta\in[\emptyset_m, ~\bar{\beta}]}s_{\theta}(z_{\theta})
$$
hence
$$
\begin{array}{lll}
s_\alpha x_\alpha s_\beta y_\beta s_\alpha (x_\alpha)^{-1} & = &\prod\limits_{\theta\in
[\emptyset_m, ~\bar{\beta}]}
s_{\alpha \cap \beta }s_{\theta}(z_{\theta}) \\
&=& \prod\limits_{\eta\in[\alpha\cap\beta, ~\beta]}s_{\eta}(z_{\eta})
\end{array}
$$
where $z_{\eta} \in NG_{n-\#\eta},$
\begin{displaymath}
z_{\nu} =\left\{\begin{array}{ll}
z_{\eta} & \text{if $\eta = \gamma_{\ast}(\eta)$} \\
1 & \text{otherwise}
\end{array}\right.
\end{displaymath}
and $s_{\alpha \cap \beta }s_{\theta} = s_\eta.$ Then $gN_n$ can be written
$$
\begin{array}{lll}
gN_n & = &\prod\limits_{\nu\in[\alpha \cap \beta, ~\beta]}s_{\nu}(z_{\nu})
\cdot
s_\beta (xy)_\beta \cdot\prod\limits_{\beta>\alpha}s_\beta y_\beta N_n.
\end{array}
$$
where $s_\eta = s_\nu.$
Thus we have shown that every product can be rewritten in the required
form modulo $N_n$, so in general, $N_n\cap D_n = NG_n\cap D_n.$ $\square$
\section{Applications and implications}
Kan introduced the notation of a $CW$-basis for a free simplicial group and used this to proved that
free simplicial groups model all connected homotopy types. The idea is that as one adds cells to the
$CW$-complex one adds new generators to the free simplicial group, but one does this within the
Moore complex so the new simplices have all but their last face at the identity element in the next
dimension down. In homotopy types where there are few such non degenerate generators or where these
generators are `generated' in a simple way then the methods behind Theorem A raise the hope of finding
a detailed presentation of the segments of the homotopy type between those dimensions in which there
are non-degenerate generators. The means of presenting this information may vary with the context, but
one set of fairly compact methods comes from the crossed algebraic techniques pioneered by J. H. C.
Whitehead in \cite{whitehead}. (Modern references for this and for more recent developments can
conveniently be found in the survey article by Baues \cite{baues3}.)
\subsection{Crossed complexes}
As an illustration we examine the impact of Theorem A on the links between simplicial groups and the
homotopy systems of Whitehead, more exactly the connected crossed complexes of Brown and Higgins
(cf. \cite{bh1} and \cite{bh2}) or the crossed chain complexes of Baues (cf. \cite{baues2} and
\cite{baues3}) as no freeness assumptions will be made here.
Let $G$ be a group, then a $G$-group is a group $H$ together with a given action of $G$ on $H,$ that is a
homomorphism from $G$ to $\text{Aut}(H).$
\begin{defn}(cf. Baues, \cite{baues3} p.22)
A crossed complex $\rho$ is a sequence
$$
\diagram
\rto^{d_4}&\rho_3\rto^{d_3}&\rho_2\rto^{d_2}&\rho_1
\enddiagram
$$
of homomorphisms between $\rho_1$-groups where $d_2$ is a crossed module and $\rho_n,$ $n\ge 3$ is
abelian and a $\pi_1$-module via the action of $\rho_1,$ where $\pi_1=\text{cokernel}(d_2).$ Moreover
$d_{n-1}d_n = 0$ for $n\ge 3.$
\end{defn}
It is known (cf. Ashley \cite{ashley}, Carrasco and Cegerra \cite{c:c} or Ehlers and Porter \cite{ep}
and the references therein) that crossed complexes correspond, via a nerve-type functor, to simplicial
groups with a `thin' structure. Each simplicial group is a Kan complex as a well known algorithm gives
a filler for any horn. A Kan complex, $K,$ is a $T$-complex if there is for each $n$ a subset $T_n$ of
$K_n,$ made up of so called `thin' elements, such that any horn has a unique thin filler and two other
more technical conditions hold (cf. Ashley \cite{ashley}). A simplicial $T$-complex which is also a
simplicial group is a group $T$-complex provided in each dimension $T_n$ is a subgroup of $K_n.$ In
this case one easily checks that $T_n$ must be $D_n$ the subgroup of $K_n$ generated by the degenerate
elements.
\begin{prop}\cite{ashley}
A simplicial group ${\bf G}$ has ${\bf NG}$ a crossed complex if and only if for each $n\geq 1,$
$NG_n\cap D_n$ is trivial.~$\square$
\end{prop}
The idea of the proof is that two $D_n$ fillers for the same horn must differ by an element of
$NG_n\cap D_n,$ so uniqueness corresponds to the simplicial group being a group $T$-complex.
The final part uses Ashley's equivalence between group $T$-complexes and crossed complexes.
Carrasco and Cegarra used this in \cite{c:c} to prove (p. 215) that a simplicial group has Moore
complex a crossed complex if and only if their pairings vanish (these are similarly defined to
those used here but are based on products rather than commutators). Similarly we have:
\begin{cor}
A necessary and sufficient condition that a simplicial group ${\bf G}$ has ${\bf NG}$ a crossed
complex is that for all $n$ and all $\alpha,\beta\in P(n),\ F_{\alpha, \beta}^{n}(x,y)$ is trivial
for all pairs $(x,y).$
\end{cor}
\begin{pf}
Since the $F_{\alpha, \beta}^{n}(x,y)$ normally generate $NG_n\cap D_n,$ this is immediate.
\end{pf}
The importance of this result is probably for the interpretation of the $F_{\alpha, \beta}^{n}$
as their vanishing has a great simplifying effect on the Moore complex.
\subsection{$\sum K(\pi,1)$}
As mentioned earlier Brown and Loday used their generalised van Kampen theorem to calculate
$\pi_3\sum K(\pi, 1)$ as $\text{Ker}(\pi\otimes\pi\longrightarrow\pi),$ the kernel of the commutator
map. Jie Wu (\cite{wu} Theorem 5.9) proves that for any group, $\pi,$ and set of generators
$\{x^{\alpha} \mid \alpha\in J\}$ for $\pi,$ then for $n\ne 1,$ $\pi_{n+2}(\sum K(\pi, 1))$ is
isomorphic to the center of the quotient group of the free product
$$
\coprod\limits_{0\leq j\leq n}(\pi)
$$
modulo the relation
$$
[y_{i_1}^{(\alpha_1)\varepsilon_1}, y_{i_2}^{(\alpha_2)\varepsilon_2},
\ldots, y_{i_t}^{(\alpha_t)\varepsilon_t}]
$$
where $\{i_1,\ldots,i_t\}=\{ -1,0,\ldots, n\}$ as sets, where $(\pi)_j$ is a copy of $\pi$
with generators $\{x_{j}^{(\alpha)}\mid \alpha\in J\}, \ \varepsilon_j= \pm 1,$
$y_{-1}^{(\alpha)}= x_0^{(\alpha)^{-1}}, \ y_{j}^{(\alpha)}= x_j^{(\alpha)}x_{j+1}^{(\alpha)^{-1}}$
for $1\leq i\leq {n-1}$ and $y_n^{(\alpha)}= x_{n}^{(\alpha)},$ and finally the commutator bracket
$[\ldots]$ runs over all the commutator bracket arrangements of weight $t$ for each $t.$
Wu's methods rely on using a construction he attributes to Carlsson \cite{carlsson}. This gives a
simplicial group $F^{\pi}(S^1)$ that has $\pi_{n+2}\sum K(\pi,1)\cong \Omega\sum K(\pi,1)\cong
\pi_{n+1}F^{G}(S^1).$ Our Theorem A above provides a link between Wu's methods and the Brown-Loday
result. We will explore this link to some extent but cannot as yet retrieve the Brown-Loday result by
purely algebraic methods. Potentially however this might yield a tensor-like description in dimension
$4$ and higher, but we will not explore that here.
Although Carlsson introduced the construction $F^G(X)$ in 1984, the construction is essentially the
same as the tensoring operation used by Quillen and others. Working in the
simplicially enriched category of
simplicial groups, there is a tensor operation defined as follows: let $K$ be a simplicial set and
$G_1, \ G_2$ simplicial groups. The simplicial group $G_1\bar{\otimes}K$ has the universal property
given by the natural isomorphism
$$
S(K, SGp(G_1, G_2))\cong SGp(G_1\bar{\otimes}K, G_2).
$$
The category of simplicial groups is also enriched over $S_{\ast},$ the category of pointed
simplicial sets. We define $G_1\bar{\wedge}K$ by
$$
S_{\ast}(K, SGp(G_1, G_2))\cong SGp(G_1\bar{\wedge}K, G_2).
$$
There is an isomorphism $F^{G}(K)\cong G\bar{\wedge}K.$ The advantage of this
approach is that it makes it clear that $\bar{\wedge}$ generalises $\wedge$
just as $\bar\otimes$ generalises $\times$
\begin{lem}
If $f: G\longrightarrow H$ is a morphism of simplicial groups, it induces
$f\bar{\wedge}K : G\bar{\wedge} K\longrightarrow H\bar{\wedge} K,$ moreover if $f$ is a weak
homotopy equivalence, so is $f\bar{\wedge}K.$
\end{lem}
\begin{pf}
As Carlsson noted, $(G\bar{\wedge}K)_n$ is
$$
\coprod\limits_{x\in K_n}(G_n)_{x}/(G_n)_{\ast}
$$
and is thus the diagonal of a bisimplicial group having
$\coprod\limits\{(G_m)_{x}\mid x\in K_n\setminus\{{\ast}\}\}$ in its $(m,n)$-position.
A simple spectral sequence argument, or direct manipulation, completes the proof.
\end{pf}
\begin{prop}
There is a natural weak homotopy equivalence
$$
\Omega\sum K(\pi,1)\simeq K(\pi,0)\bar{\wedge}S^{1}
$$
where $K(\pi,0)$ is the constant simplicial group with value, $\pi,$ $S^1$ is the simplicial
$1$-sphere and $\sum$ is reduced suspension.
\end{prop}
\begin{pf}
As Kan's loop group functor models $\Omega$ and $K\wedge S^1$ the suspension,
$$
\Omega(\sum K(\pi, 1))\simeq G(K(\pi,1)\wedge S^1)
$$
then the adjunction between $G$ and the classifying space functor $\bar{W}$
gives for $K$, $ L$, arbitrary pointed simplicial sets, and ${\bf H}$ an arbitrary
simplicial group, the natural isomorphisms
$$
\begin{array}{llll}
SGp(G(K)\bar{\wedge}L, {\bf H})&\cong& S_{\ast}(L, SGp(G(K), {\bf H}))\\
& \cong& S_{\ast}(L, S_{\ast}(K, \bar{W}{\bf H})) \\
& \cong& S(K\wedge L, \bar{W}{\bf H})\\
& \cong & SGp(G(K\wedge L), {\bf H})
\end{array}
$$
thus $G(K)\bar{\wedge}L\cong G(K\wedge L).$ As Curtis notes (\cite{curtis} p. 137)
$\bar{W}K(\pi, 0)$ is a minimal complex for $K(\pi,1)$ so taking $K = \bar{W}K(\pi,0)$
we get $\Omega\sum K(\pi, 1)$ has as model $G(\bar{W}(K(\pi, 0))\wedge S^1)$ and hence
$G(\bar{W}(K(\pi, 0)))\bar{\wedge}S^1.$ By Lemma, 4.4 given the weak homotopy equivalence
$G\bar{W}(K(\pi, 0))\longrightarrow K(\pi,0),$ the result follows. \end{pf}
This implies that, like Jie Wu \cite{wu}, we can take a simple model for $\Omega\sum K(\pi, 1).$
First we introduce notation for $S^1.$ We write $S_0^1 =\ast, \ S_1^1 =\{\sigma,\ast\}, \
S_2^1 =\{ x_0, x_1, \ast \}$ 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
$x_i =s_n\ldots s_{i+1}s_{i-1}\dots s_0\sigma,$ $0\leq i \leq n.$
For simplicity we write $G= K(\pi,0)$ and make no distinction between simplicies in
different dimensions unless confusion might arise. This then gives
$$
\begin{array}{llll}
(G\bar\wedge S^1)_0 &=& 1, \qquad\text{the trivial group}\\
(G\bar\wedge S^1)_1 &\cong& \pi, \\
(G\bar\wedge S^1)_1 &\cong&\pi\ast\pi, \qquad\text{the free product of two copies of $\pi$}\\
(G\bar\wedge S^1)_3 &\cong&\pi\ast\pi\ast\pi \qquad\text{and so on}.
\end{array}
$$
We write $g\bar{\wedge}x$ for the $x$-indexed copy of $g\in \pi$ in the coproduct
$\coprod\{(\pi)_x : x\in S_n^1\setminus\{{\ast}\}\}$ so the only relations we have are of the form
$$
(g{g'}\bar{\wedge} x) = (g\bar{\wedge} x)({g'}\bar{\wedge} x).
$$
We next analyse $N(G\bar{\wedge}S^1)$ in low dimensions. For simplicity we will write
${\bf H}$ instead of $G\bar{\wedge}S^1.$ Of course $NH_0 =1,\ NH_1 = \pi.$
By Theorem A, $NH_2$ is generated by all $F_{(0)(1)}(g\bar{\wedge}\sigma, h\bar{\wedge}\sigma),$
$g,\ h \in \pi:$
$$
F_{(0)(1)}(g\bar{\wedge}\sigma, h\bar{\wedge}\sigma)=
(g\bar{\wedge}x_1)(h\bar{\wedge}x_0)(g^{-1}\bar{\wedge}x_1)(gh^{-1}g^{-1}\bar{\wedge}x_0).
$$
In fact although Theorem A gives these as normal generators, it is clear that these
are generators since conjugates of them are expressible as product of other
terms of the same form. For instance
$$
{}^{k\bar{\wedge}x_1}F_{(0)(1)}(g\bar{\wedge}\sigma, h\bar{\wedge}\sigma)=
F_{(0)(1)}(kg\bar{\wedge}\sigma, h\bar{\wedge}\sigma)
F_{(0)(1)}(kg\bar{\wedge}\sigma, ghg^{-1}\bar{\wedge}\sigma)
$$
and a similar expression can be found for conjugation by $k\bar{\wedge}x_0.$
Brown and Loday \cite{bl1} calculated $\pi_3(\sum K(\pi, 1))$ using a van Kampen
theorem for cat$^2$-groups. This led to an expression for this group as being isomorphic to
$$
J_2(\pi) = \mbox{Ker}(\kappa: \pi\otimes\pi\longrightarrow\pi).
$$
We refer to the paper \cite{bjr} Brown, Johnson and Robertson for some details on the non-abelian tensor
product of groups and to Ellis \cite{ellis2} for more on the representation of homotopy types by
crossed squares and cat$^2$-groups. Here it will suffice to say that if $G,$ $H$ are groups that act
on themselves by conjugation and on each other in such a way that
$$
{}^{{}^{g}h}{g'}={}^{ghg^{-1}}{g'}\hspace{2cm}{}^{{}^{h}g}{h'}={}^{hgh^{-1}}{h'}
$$
(see \cite{bjr}), the tensor product $G\otimes H$ is the group generated by symbols $g\otimes h$
with relations
$$
\begin{array}{lcr}
g{g'}\otimes h &=& ({}^{g}{g'}\otimes {}^{g}h)(g\otimes h)\cr
g\otimes h{h'} &=& (g\otimes h)\otimes({}^{h}g\otimes {}^{h}{h'})
\end{array}
$$
for all $g,{g'}\in G,$ $h,{h'}\in H.$ We will need this only when $G = H.$, then there is a map
$$
\kappa: G\otimes G\longrightarrow G
$$
given by $\kappa(g\otimes h)=[g,~h].$ This gives a homomorphism since the above relations are abstract
versions of the usual commutator identities.
Using a combination of Ellis's work in \cite{ellis2} and the second author's description of the
crossed $n$-cube associated to a simplicial group in \cite{porter}, it is clear that the expression
for the tensor product should be closely linked to one for $NH_2/d_3NH_3$ and with that in mind we set
for $g,~h\in\pi$
$$
g\bar{\otimes}h = [g\bar{\wedge}x_0,~(h\bar{\wedge}x_0)(h^{-1}\bar{\wedge}x_1)]~d_3NH_3.
$$
This `mysterious' formula will be more fully explained in another paper where the relationship
with crossed squares and Ellis's work will be given in detail. For the moment the reader is asked
merely to accept the left hand side as a shorthand for the right hand side.
\begin{prop}
The symbols $g\bar{\otimes} h$ generate $NH_2/d_3NH_3$ and satisfy the tensor product relations above.
The boundary homomorphism
$$
{d_2'}:\frac{NH_2}{d_3NH_3}\longrightarrow NH_1
$$
sends $g\bar{\otimes} h$ to $[g,~h]\bar{\wedge}\sigma$ in $NH_1.$
\end{prop}
\begin{pf}
Direct calculation shows
$$
g\bar{\otimes} h = F_{(0)(1)}(h\bar{\wedge}\sigma, g\bar{\wedge}\sigma)~d_3NH_3
$$
which clearly generate $NH_2/d_3NH_3$ by Theorem A and the commutator above.
The tensor product relations are again directly verifiable (eg. using the description of the
$h$-map given in \cite{porter}), and it is immediate that
$$
d_2[g\bar{\wedge}x_0,~(h\bar{\wedge}x_0)(h^{-1}\bar{\wedge}x_1)]= [g,~h].
$$\hfill\end{pf}
The next term in the Moore complex $NH_3$ has 6 different types of generator and so there are 6
$\underline{\mbox{known}}$ types of relation in $NH_2/d_3NH_3.$ Presumably it is possible to give
a direct proof that
$$
\frac{NH_2}{d_3NH_3}\cong\pi\otimes\pi
$$
using these relation, but we have so far not managed to find one. A closer analysis of the relationship
of the Peiffer pairings with the structure of crossed squares gives this isomorphism by a universal
argument, but requires other techniques and so will be postponed to a later paper. Knowledge of the 6
types of generator for $NH_3$ hopefully will allows a detailed calculation of $NH_3/d_4NH_4$ in a similar
manner.
|
\part{#1}
\pagestyle{myheadings}
\markright{Eaves, Ph.D\hfill1999\hfill\ Page }
}
\urlstyle{sf}
\newcommand{\myURL}[1]{
\href{#1}{\url{#1}}
}
\newcommand{\myRef}[1]{
\S\ref{#1}
}
\newcommand{\textit{etc.}}{\textit{etc.}}
\newcommand{\textit{e.g.}\,}{\textit{e.g.}\,}
\newcommand{\textit{i.e.}\,}{\textit{i.e.}\,}
\newcommand{\textit{vs.}\,}{\textit{vs.}\,}
\newcommand{\textit{et al.}\,}{\textit{et al.}\,}
\newcommand{\textit{op. cit.}\,}{\textit{op. cit.}\,}
\newcommand{\textit{viz.}\,}{\textit{viz.}\,}
\newcommand{\textit{NB.}\,}{\textit{NB.}\,}
\newcommand{\textit{vs.}\,}{\textit{vs.}\,}
\newcommand{\emph{Aidan}\,}{\emph{Aidan}\,}
\newcommand{\emph{CoBase}\,}{\emph{CoBase}\,}
\newcommand{\emph{Java}\,}{\emph{Java}\,}
\newcommand{\emph{Jigsaw}\,}{\emph{Jigsaw}\,}
\newcommand{\emph{PostgreSQL}\,}{\emph{PostgreSQL}\,}
\newcommand{\mySrc}[1]{\textsf{#1}}
\newlength{\facewd} \newlength{\faceht}
\newcommand{\markOver}[1]{
\settowidth{\facewd}{#1}
\settoheight{\faceht}{#1}
\raisebox{\faceht}[0pt]{
\makebox[0pt][l]{
\hspace{.15\facewd}$\blacktriangledown$
}
}
#1
}
\newcommand{\markUnder}[1]{
\settowidth{\facewd}{#1}
\settoheight{\faceht}{#1}
\raisebox{-\faceht}[0pt]{
\makebox[0pt][l]{
\hspace{.15\facewd}$\blacktriangledown$
}
}
#1
}
\newcommand{\myImport}[1]{
\markOver{\mySrc{#1}}
}
\newcommand{\myExport}[1]{
\markUnder{\mySrc{#1}}
}
\newcommand{\myAttr}[1]{
\mySrc{#1}
}
\newcommand{\myEnv}[1]{
\texttt{#1}
}
\newenvironment{myList}[1]{
\begin{list}{\#1{bean}}{\usecounter{bean}}
\setlength{\rightmargin}{\leftmargin}
}
{
\end{list}
}
\theoremstyle{plain}
\newtheorem{theorem}{Theorem}[section]
\newtheorem{lemma}{Lemma}[section]
\newtheorem{corollary}{Corollary}[section]
\theoremstyle{definition}
\newtheorem{proposition}{Proposition}[section]
\newtheorem{scenario}{Scenario}[section]
\newtheorem{condition}{Condition}[section]
\newtheorem{definition}{Definition}[section]
\newtheorem{notation}{Notation}[section]
\newtheorem{remark}{Remark}[section]
\newtheorem{mrule}{Rule}[section]
\newtheorem{axiom}{Axiom}[section]
\newenvironment{mySubequations}{
\begin{subequations}
\renewcommand{\theequation}{\theparentequation \roman{equation}}
}
{
\end{subequations}
}
\newcommand{\droptext}[1]{\ensuremath{
\text{\ #1\ }}
}
\newcommand{\eqsubst}[2]{\ensuremath{
\left[ #1 \vert #2 \right]
}
}
\newcommand{\identity}[1]{\ensuremath{
\langle #1 \rangle
}
}
\newcommand{\pubKey}[1]{\ensuremath{
+{#1}
}
}
\newcommand{\privKey}[1]{\ensuremath{
-{#1}
}
}
\newcommand{\encrypted}[2]{\ensuremath{
\{{#1}\}_{#2}
}
}
\newcommand{\decrypted}[2]{\ensuremath{
\{{#1}\}_{#2}^{-1}
}
}
\newcommand{\hashed}[1]{\ensuremath{
H({#1})
}
}
\newcommand{\func}[1]{\ensuremath{
F({#1})
}
}
\newcommand{\fresh}[1]{\ensuremath{
\mathrel{\# \left( {#1} \right) }
}
}
\newcommand{\recognizable}[1]{\ensuremath{
\mathrel{\phi \left( {#1} \right) }
}
}
\newcommand{\sees}{\ensuremath{
\mathrel{\vartriangleleft}
}
}
\newcommand{\holds}{\ensuremath{
\ni
}
}
\newcommand{\said}{\ensuremath{
\mathrel{\mid\sim}
}
}
\newcommand{\believes}{\ensuremath{
\mathrel{\mid\equiv}
}
}
\newcommand{\controls}{\ensuremath{
\mathrel{\mid\Rightarrow}
}
}
\newcommand{\secret}[1]{\ensuremath{
\mathrel{\stackrel{#1}{\leftrightarrow}}
}
}
\newcommand{\means}{\ensuremath{
\rightsquigarrow
}
}
\newcommand{\project}[1]{\ensuremath{
\mathop{\stackrel{#1}{\rightarrow}}
}
}
\newcommand{\rvec}[1]{\ensuremath{
\overset{\leftarrow}{#1}
}
}
\newcommand{\controlEntity}[1]{\ensuremath{
\operatornamewithlimits{control}_{#1}
}
}
\newcommand{\createEntity}[1]{\ensuremath{
\operatornamewithlimits{create}_{\text{with\ } #1}
}
}
\newcommand{\receiveRight}[2]{\ensuremath{
\operatornamewithlimits{receive}_{#1 \droptext{from} #2}
}
}
\newcommand{\grantRight}[2]{\ensuremath{
\operatornamewithlimits{grant}_{#1 \droptext{to} #2}
}
}
\newcommand{\takeRight}[2]{\ensuremath{
\operatornamewithlimits{take}_{#1 \droptext{from} #2}
}
}
\newcommand{\typeOf}{\ensuremath{
\operatorname{type}
}
}
\newcommand{\instanceOf}{\ensuremath{
\operatorname{instance}
}
}
\newcommand{\accrete}{\ensuremath{
\operatorname{accrete}
}
}
\newcommand{\retract}{\ensuremath{
\operatorname{retract}
}
}
\newcommand{\cost}{\ensuremath{
\operatorname{cost}
}
}
\newcommand{\bigsqcap}{\ensuremath{
\sqcap
}
}
\newcommand{\field}[1]{\ensuremath{\Bbb{#1}}}
\newcommand{\ensuremath{\triangleq}}{\ensuremath{\triangleq}}
\newcommand{\powerSet}[1]{\ensuremath{
\mathscr{P}\left(#1\right)
}
}
\newcommand{\rightarrowdotted}[1]{\ensuremath{
\mathrel{\xrightarrow[#1]{}}}
}
\newcommand{\truthSet}[2]{\ensuremath{
{ \parallel #1 \parallel }^{#2}
}
}
\newcommand{ \ensuremath{ \mathrel{R} } }{ \ensuremath{ \mathrel{R} } }
\newcommand{ \ensuremath{ \operatorname{Supp} } }{ \ensuremath{ \operatorname{Supp} } }
\newcommand{ \ensuremath{ \operatorname{Ran} } }{ \ensuremath{ \operatorname{Ran} } }
\newcommand{ \ensuremath{ \operatorname{Dom} } }{ \ensuremath{ \operatorname{Dom} } }
\newcommand{ \ensuremath{ \operatorname{Path} } }{ \ensuremath{ \operatorname{Path} } }
\newcommand{\propn}[1]{ \ensuremath{\mathbb{P}_{#1}}}
\newcommand{\propnSet}[1]{ \ensuremath{P_{#1}}}
\newcommand{\myTheoremWithin}[1]{
\ensuremath{ \mathrel {
\underset{#1}{\vdash}}
}
}
\newcommand{\ensuremath{\mathscr{L}}}{\ensuremath{\mathscr{L}}}
\newcommand{\myClassName}[1] {\ensuremath{
\boldsymbol{#1}
}
}
\newcommand{\myModelName}[1] {\ensuremath{
\mathscr{#1}
}
}
\newcommand{\myLogicName}[1] {\ensuremath{
\mathscr{#1}
}
}
\newcommand{\myModel}{\ensuremath{
\myModelName{U} =
\langle
{\mathcal W}, \dots , {\mathcal P }
\rangle
}
}
\newcommand{\myModelExtended}[1]{\ensuremath{
\myModelName{U} =
\langle
{\mathcal{W}}, #1 , {\mathcal{P}}
\rangle
}
}
\newcommand{\myModelForByAt}[2]{
\ensuremath{ \mathrel {
\overset{#1}{\underset{#2}{\models}}
}
}
}
\newcommand{\myModelForAt}[1]{ \myModelForByAt{\myModelName{U}}{#1} }
\newcommand{ \myModelForByAt{\myModelName{U}}{x} }{ \myModelForByAt{\myModelName{U}}{x} }
\newcommand{ \myModelForByAt{\myModelName{U}}{} }{ \myModelForByAt{\myModelName{U}}{} }
\newcommand{ \myModelForByAt{}{\myClassName{C}} }{ \myModelForByAt{}{\myClassName{C}} }
\newcommand{ \myModelForByAt{}{\myLogic} }{ \myModelForByAt{}{\ensuremath{\mathscr{L}}} }
\newcommand{\ensuremath{\mathop{\square}}}{\ensuremath{\mathop{\square}}}
\newcommand{\ensuremath{\mathop{\lozenge}}}{\ensuremath{\mathop{\lozenge}}}
\newcommand{\isModal}{\ensuremath{
\mathop{\left[ \ensuremath{\mathop{\lozenge}} \mid \ensuremath{\mathop{\square}} \right]}
}
}
\newcommand{\slaveLogic}{\ensuremath{
\myModelName{U_{\top}}
}
}
\newcommand{\modalCalculus}[1]{ \textbf{#1} }
\newcommand{\ensuremath{\operatorname{O}}}{\ensuremath{\operatorname{O}}}
\newcommand{\ensuremath{\operatorname{P}}}{\ensuremath{\operatorname{P}}}
\newcommand{\ensuremath{\operatorname{F}}}{\ensuremath{\operatorname{F}}}
\newcommand{\take}{\ensuremath{
\operatorname{take}
}
}
\newcommand{\seen}{\ensuremath{
\operatorname{seen}
}
}
\newcommand{\punish}{\ensuremath{
\operatorname{punish}
}
}
\newcommand{\worth}{\ensuremath{
\operatorname{worth}
}
}
\newcommand{\challenged}{\ensuremath{
\operatorname{challenged}
}
}
\newcommand{\myVec}[2]{\ensuremath{
\Vec{#1}^{\langle #2 \rangle }
}
}
\newcommand{\myMat}[2]{\ensuremath{
\Bar{#1}^{\langle #2 \rangle }
}
}
\newcommand{\dset}{\ensuremath{
\{ 1, 0, -1 \}
}
}
\newcommand{\sign}{\ensuremath{
\operatorname{\textbf{s}}
}
}
\newcommand{\promote}{\ensuremath{
\operatorname{promote}
}
}
\newcommand{\demote}{\ensuremath{
\operatorname{demote}
}
}
\newcommand{\sincere}{\ensuremath{
\operatorname{sincere}
}
}
\newcommand{\sophis}{\ensuremath{
\operatorname{sophisticated}
}
}
\begin{document}
\title{Transport Level Security: a proof using the Gong--Needham--Yahalom
Logic}
\author{Walter D Eaves}
\date{\today}
\maketitle
\begin{abstract}
This paper provides a proof of the proposed Internet standard Transport
Level Security protocol using the Gong--Needham--Yahalom logic. It is
intended as a teaching aid and hopes to show to students: the potency of
a formal method for protocol design; some of the subtleties of
authenticating parties on a network where all messages can be
intercepted; the design of what should be a widely accepted standard.
\end{abstract}
\section{Transport Level Security Protocol}
\label{sec:auth}
This section provides an insight into the workings of the next
generation of authentication protocol: the Transport Level Security
Protocol version 1.0\cite{draft:tls}, the successor to the Secure
Sockets Layer\cite{draft:ssl}. To do this, the Gong--Needham--Yahalom,
GNY, logic \cite{GoNeYa90} is introduced which is a formal method for
proving the safety of a cryptographically-based protocol. It is
described at length in appendix \ref{cha:protocols}. When working
through protocols the relevant rule of inference will be stated and
will refer to those in the appendix.
The \emph{Transport Level Security handshake protocol}\cite{draft:tls},
TLS, has an unknown heritage, but it has a great deal of similarity to that
described in \cite{Denning:1981:TKD}. It is predicated on the existence of
readily available public keys: TLS's predecessor made use of X.509
certificates, see \cite{CCITTConsult88b}, issued by a Certification
Authority, CA, an example of which is \textit{Thawte}\cite{ca:thawte}. A
discussion of the limitations of certificate technology can be found in
R\"oscheisen's on--line paper \cite{Rosche95}.
TLS has three sub--protocols:
\begin{itemize}
\item Server anonymous
\item Server named, client anonymous
\item Server named, client named
\end{itemize}
These differ by who is required to send their X.509 certificates, the key
exchange protocol is different only when the client is named and thus has a
public--key that can be used. The messages are shown in figure
\ref{fig:auth-2} sent during a run of the protocol are more or less the
same for all sub--protocols. As can be seen, no key issuing server is
needed.
The TLS is a more complicated protocol than the \textit{Kerberos} which is
described in the appendix \myRef{sec:kerb}, before looking at TLS's
protocol proof, it might be best to examine \textit{Kerberos's}. Also,
there are some more examples of other authentication protocols being
investigated and found lacking\cite{Lowe96b}.
A protocol proof has three stages:
\begin{itemize}
\item Message Analysis
\item Pre--conditions Analysis
\item Belief deductions for each message
\end{itemize}
Message analysis involves formalizing the content of messages so that they
contain just keys and identifiers. Pre--conditions analysis formalizes what
the parties to the protocol assume about the state of keys before
undertaking the protocol run. Belief deductions analyzes how each party can
deduce new beliefs when it receives a new message.
\section{Messages for the Named Server Protocol}
There are six message exchanges. There is a provision for more, to settle
which cryptographic implementations to use and for the client to provide a
certificate, but this is the basic protocol for a named server to an
anonymous client.
\begin{figure}[htbp]
\begin{center}
\includegraphics{auth-2.eps}
\caption{TLS protocol: Messages}
\label{fig:auth-2}
\end{center}
\end{figure}
\newcommand{\secretKey}{\ensuremath{
\func{(N_A, T_A, N_B, T_B), {N'}_A}
}
}
\newcommand{\certificate}{\ensuremath{
\encrypted{\pubKey{K_B}, \identity{B}, \identity{C}}{\privKey{K_C}}
}
}
\paragraph{Messages}
\begin{align}
A \rightarrow B & \colon
(N_A, T_A)
\label{eq:tls:m1} \tag{$M_1$} \\
%
B \rightarrow A & \colon
(N_B, T_B)
\label{eq:tls:m2} \tag{$M_2$} \\
%
B \rightarrow A & \colon
\certificate
\label{eq:tls:m3} \tag{$M_3$} \\
%
A \rightarrow B & \colon
\encrypted{{N'}_A}{\pubKey{K_B}}
\label{eq:tls:m4} \tag{$M_4$} \\
%
B \rightarrow A & \colon
\encrypted{H(K_{AB}, AB_5, (M_1, M_2, M_3, M_4))}{K_{AB}}
\label{eq:tls:m5} \tag{$M_5$} \\
%
A \rightarrow B & \colon
\encrypted{H(K_{AB}, AB_6, (M_1, M_2, M_3, M_4))}{K_{AB}}
\label{eq:tls:m6} \tag{$M_6$} \\
\intertext{where}
K_{AB} & = \secretKey \notag \\
\intertext{and, the behaviour of the public and private keys is:}
\encrypted{\encrypted{X}{\privKey{K}}}{\pubKey{K}} & = X \notag \\
\encrypted{\encrypted{X}{\pubKey{K}}}{\privKey{K}} & = X \notag
\end{align}
and
\begin{equation*}
\begin{aligned}
AB_5 \ensuremath{\triangleq} \droptext{``server finished''}
\end{aligned}
\droptext{,}
\begin{aligned}
AB_6 \ensuremath{\triangleq} \droptext{``client finished''}
\end{aligned}
\end{equation*}
The messages can be summarized as follows:
\begin{description}
\item[\ref{eq:tls:m1}] $A$ sends a timestamp and a nonce to $B$.
\item[\ref{eq:tls:m2}] $B$ sends another timestamp and another nonce to $A$.
\item[\ref{eq:tls:m3}] $B$ sends its certificate signed by the
certification authority; it contains $\pubKey{K_B}$, $B$'s public--key.
\item[\ref{eq:tls:m4}] $A$ returns the ``pre--master secret'' ${N'}_A$
encrypted under $\pubKey{K_B}$.
\item[\ref{eq:tls:m5}] $B$ sends a hash of the session key, a tag
indicating the protocol stage $AB_5$\footnote{Actually all stages of the
protocol are marked with a stage identifier, but it is not necessary to
consider all of them.} and all preceding messages exchanged to $A$.
\item[\ref{eq:tls:m6}] $A$ sends a hash of the session key, a tag
indicating the protocol stage $AB_6$ key and all preceding messages to
$B$.
\end{description}
\section{Pre--Conditions}
\begin{mySubequations}
\paragraph{Certificates}
\begin{enumerate}
\item Some Expectations
$A$ expects to use $C$ as the certification authority and expects to use
$B$, so
\begin{equation}
\label{eq:tls:c:0}
\begin{aligned}
A & \holds \identity{C}
\end{aligned}
\quad
\begin{aligned}
A & \holds \identity{B}
\end{aligned}
\end{equation}
\item Using Them
The role of the unseen certification authority, $C$, is pivotal. Even
though $C$ has used the private key $\privKey{K_C}$ to create the
certificate, this key is not used again.
\begin{equation}
\begin{aligned}
B \believes A \believes \project{\pubKey{K_C}} C
\end{aligned}
\text{,} \quad
\begin{aligned}
A \believes \project{\pubKey{K_C}} C
\end{aligned}
\label{eq:tls:c:1}
\end{equation}
The two parties rely on the public--key being available. $A$ must be able
to get $K_C$:
\begin{equation}
\begin{aligned}
A \holds \pubKey{K_C}
\end{aligned}
\text{,} \quad
\begin{aligned}
B \believes A \holds \pubKey{K_C}
\end{aligned}
\label{eq:tls:c:2}
\end{equation}
\item Trusting in them
The pre--conditions regarding $B$'s certificate are as follows. $A$ and
$B$ both trust $C$ to deliver the correct identity with the public key.
\begin{equation}
\begin{aligned}
A \believes C \controls ( \pubKey{K_P}, \identity{P} )
\end{aligned}
\text{,} \quad
\begin{aligned}
B \believes C \controls ( \pubKey{K_P}, \identity{P} )
\end{aligned}
\label{eq:tls:c:3}
\end{equation}
\item Meaning of the contents
For $A$ the assumption underlying a certificate is that the public--key is
the public--key of the named party.
\begin{equation}
A \believes ( \pubKey{K_P}, \identity{P} ) \means
C \believes \project{\pubKey{K_P}} P
\label{eq:tls:c:4}
\end{equation}
And $A$ believes $C$ when $C$ names a key:
\begin{equation}
\label{eq:tls:c:5}
A \believes C \controls \project{\pubKey{K_P}} P
\end{equation}
\end{enumerate}
\paragraph{System Capabilities}
\begin{enumerate}
\item $B$ believes that $A$ can generate a nonce and keep it secret to pass
it on as the pre--master secret.
\begin{equation}
\begin{aligned}
B \believes A \controls \fresh{X}
\end{aligned}
\text{,} \quad
\begin{aligned}
B \believes A \secret{X} A
\end{aligned}
\label{eq:tls:c:6}
\end{equation}
\item $A$ and $B$ have both assumed that the other can generate the master
secret if presented with the components.
\begin{equation}
\begin{aligned}
A \believes B \controls \func{X,Y}
\end{aligned}
\text{,} \quad
\begin{aligned}
B \believes A \controls \func{X,Y}
\end{aligned}
\label{eq:tls:c:7}
\end{equation}
and, of course, they do
\begin{equation}
\begin{aligned}
B \controls \func{X,Y}
\end{aligned}
\text{,} \quad
\begin{aligned}
A \controls \func{X,Y}
\end{aligned}
\label{eq:tls:c:8}
\end{equation}
\item $B$ has a private--key and holds his own certificate.
\begin{equation}
\begin{aligned}
B \holds \privKey{K_B}
\end{aligned}
\text{,} \quad
\begin{aligned}
B \holds \certificate
\end{aligned}
\label{eq:tls:c:9}
\end{equation}
Note that the format of a certificate does includes a statement of the
identity of $C$. Although not used in this protocol it is an important
part of it since it allows the public--key of $C$ to be checked.
\end{enumerate}
\end{mySubequations}
\section{Belief Deductions Analysis}
\begin{mySubequations}
\begin{enumerate}
\item Messages \ref{eq:tls:m1} received by $B$ and \ref{eq:tls:m2} received
by $A$
These nonce and timestamp exchanges are important, because they are used
in the generation of the key. The vindication is the appearance of the
time--stamp, which is definitely fresh, and the the rule
\eqref{eq:gny:f1} freshens the nonces.
\begin{equation}
\begin{aligned}
A & \holds M_1, M_2 \droptext{By \eqref{eq:gny:p1}} \\
A & \holds N_B, T_B, N_A, T_A \\
A & \believes \fresh{N_B}
\end{aligned}
\droptext{and}
\begin{aligned}
B & \holds M_1, M_2 \droptext{By \eqref{eq:gny:p1}} \\
B & \holds N_A, T_A, N_B, T_B \\
B & \believes \fresh{N_A}
\end{aligned}
\label{eq:tls:a:1}
\end{equation}
\item Message \ref{eq:tls:m3} received by $A$
$A$ receives the certificate from, presumably, $B$. By
\eqref{eq:tls:c:1}, \eqref{eq:tls:c:2} and \eqref{eq:gny:i4}, $A$ now
has:
\begin{equation*}
A \believes C \said \certificate
\end{equation*}
and can decrypt the contents to discover:
\begin{equation*}
C \said ( \pubKey{K_B}, \identity{B} )
\end{equation*}
By \eqref{eq:tls:c:3} and \eqref{eq:tls:c:4}
\begin{equation*}
A \believes C \believes \project{\pubKey{K_B}} B
\end{equation*}
By \eqref{eq:tls:c:5} and \eqref{eq:gny:j1}
\begin{equation*}
A \believes \project{\pubKey{K_B}} B
\end{equation*}
And of course
\begin{equation}
\begin{split}
A \holds M_3 & \droptext{By \eqref{eq:gny:p1}.} \\
A \holds M_4 & \droptext{Because $A$ creates it.}
\end{split}
\label{eq:tls:a:3}
\end{equation}
Notice that $A$ does not know that the sender of this message was $B$.
\item Message \ref{eq:tls:m4} received by $B$
$B$ gains knowledge of the following:
\begin{equation}
\begin{split}
B \holds M_3 & \droptext{Since it sent it} \\
B \holds M_4 & \droptext{By \eqref{eq:gny:p1}}
\end{split}
\label{eq:tls:a4}
\end{equation}
From which by \eqref{eq:gny:i2} where $\identity{S}$ is ${N'}_A$.
\begin{align*}
B & \sees {N'}_A \\
B & \holds {N'}_A
\end{align*}
This is something of an innovation: ${N'}_A$ has not been established as
a shared secret by the conventional method of passing it along a channel
secured by a long--term key or comparing it to a pre--stored hash, but it
\emph{will} be established as a secret by the subsequent correct
operation of the protocol.
Since ${N'}_A$ came under cover of the public key and it will be later
identified as unique to the sender, it is a shared secret, so by
\eqref{eq:tls:c:6}.
\begin{align*}
B & \believes A \secret{{N'}_A} B \\
B & \believes \fresh{{N'}_A}
\end{align*}
Now by \eqref{eq:tls:c:8} and \eqref{eq:tls:a:1}
\begin{align*}
B & \holds K_{AB} \\
B & \believes \fresh{K_{AB}} \\
\because B & \holds N_A, N_B, T_A, T_B
\end{align*}
$B$ can now construct the response message:
\begin{align*}
B & \holds M_1, M_2, M_3, M_4 \\
B & \holds H(K_{AB}, AB_5, (M_1, M_2, M_3, M_4))
\end{align*}
\item Message \ref{eq:tls:m5} received by $A$
$A$ now receives a message from $B$ which can only be understood
\emph{correctly}, if both $A$ and $B$ have agreed upon $K_{AB}$.
$A$ performs some pre--calculations:
\begin{align*}
A & \holds K_{AB} \\
\because A & \holds {N'}_A, N_A, N_B, T_A, T_B \\
A & \holds H(K_{AB}, AB_5, (M_1, M_2, M_3, M_4))
\end{align*}
Because $A$ has been collecting the messages as well \eqref{eq:tls:c:8}
and holding all previous messages \eqref{eq:tls:a:1} and
\eqref{eq:tls:a:3}.
By \eqref{eq:gny:i1} $A$ can decrypt the message
\begin{align*}
A \sees & H(K_{AB}, AB_5, (M_1, M_2, M_3, M_4)) \\
\therefore & \\
A \believes & H(K_{AB}, AB_5, (M_1, M_2, M_3, M_4)) \\
& \means B \holds H(K_{AB}, AB_5, (M_1, M_2, M_3, M_4))
\end{align*}
$A$ can now make a series of justified conclusions, by \eqref{eq:gny:i3}
\begin{align*}
A & \believes B \holds {N'}_B, K_{AB} \\
A & \believes B \holds M_1, M_2, M_3, M_4 \\
A & \believes B \holds N_A, T_A, N_B, T_B
\end{align*}
$A$ can now validate the identity of the other party:
\begin{equation}
\begin{aligned}
A & \believes B \said (N_B, T_B) \\
A & \believes B \believes A \said (N_A, T_A) \\
A & \believes B \believes A \secret{K_{AB}} B \\
A & \believes B \believes \fresh{K_{AB}}
\end{aligned}
\text{\quad}
\begin{aligned}
A & \believes B \sees M_1 \\
A & \believes B \said M_2 \\
A & \believes B \said M_3 \\
A & \believes B \sees M_4
\end{aligned}
\label{eq:tls:a:5}
\end{equation}
It should be clear now why the key $K_{AB}$ is hashed into the hash
signature $H(K_{AB}, AB_5, (M_1, M_2, M_3, M_4))$. A hash is only
validated by inclusion of a secret and a nonce, see \eqref{eq:gny:i3},
the key $K_{AB}$ is both.
\item Message \ref{eq:tls:m6} received by $B$
By a similar argument to that used for $A$, it is clear:
\begin{equation}
\begin{aligned}
B & \believes A \said {N'}_A \\
B & \believes A \believes B \said (N_B, T_B) \\
B & \believes A \believes A \secret{K_{AB}} B \\
B & \believes A \believes \fresh{K_{AB}}
\end{aligned}
\text{\quad}
\begin{aligned}
B & \believes A \said M_1 \\
B & \believes A \sees M_2 \\
B & \believes A \sees M_3 \\
B & \believes A \said M_4
\end{aligned}
\end{equation}
Since $A$ could only generate this message if in possession of $K_{AB}$,
$B$ can deduce that $A$ is the party with whom it shares the key and the
whole protocol run is current.
\end{enumerate}
\end{mySubequations}
\section{Summary: Innovations and Possible Attacks}
\paragraph{Summary}
The Transport Level Security handshake protocol is quite ingenious: it lets
$A$ send a random message under a public key which is used as an
identifying secret shared by the parties before it has been established as
such. A challenge and response protocol, the challenge is issued in
plain--text and the response returns it as cipher--text, so that the
challenger can verify that the responder knows the shared session key. This
protocol is effectively a challenge and response protocol with the
generated session key, which is created from the secret sent under the
public--key.
The protocol is also exemplary in its use of stage identifiers and hash
digests. The stage identifiers change the hash digest between messages
\ref{eq:tls:m5} and \ref{eq:tls:m6}. The hash digests validate the
whole protocol run.
\paragraph{Attacks}
The critical point of the protocol is the transmission of the public--key
with which the client should respond with a nonce encrypted under it. The
man-in-the-middle attack is well known here: all that need be done is to
intercept \ref{eq:tls:m3} and substitute a bogus certificate. The fraud
then hinges upons the expectations of $A$, \eqref{eq:tls:c:0}.
\begin{enumerate}
\item Impersonate $C$ and $B$
If $A$ is not expecting to use $C$ or $B$ then the attacker, $M$, can
substitute a different certificate for another service: $D$.
\item Impersonate $B$
If $A$ is not expecting to use $B$ but is expecting
a certificate issued by $C$ then $M$ can create a service $D$ and attempt
to have $C$ issue a certificate for it.
\end{enumerate}
It is quite easy to provide a service that looks like $B$ and appears to be
at the address of $B$ this is rather more difficult with a certification
authority because the public certification scheme proposed in
\cite{OSI:dir} is based upon the following:
\begin{itemize}
\item Certification authorities are well--known in that their addresses and
public--keys can be obtained from many sources.
\item Certificates contain lists of certification authorities which allow a
client to match known certification authorities to those found in the
certificates.
\end{itemize}
Certification authorities currently do nothing other than provide
certificates, so all a client obtains from a certificate is some
accountability. If defrauded the client can attempt to locate the server
who perpetrated the fraud.
\paragraph{Another Useful Feature}
One of the provisions of TLS is to allow the server to pass to the client
another key to use in place of the certificate key. This may be necessary
for any of the following reasons:
\begin{itemize}
\item The client lacks an implementation to encrypt with the server's
public--key.
\item The server does not wish to use its public--key.
\item Restrictions on key size require that a smaller or larger key must be
used.
\end{itemize}
The client would receive a different public--key but that must be signed
under the certificate key for the client to have any faith in it. If the
client had chosen to use the alternative key because it lacked an
implementation, it would still need to decrypt under the certification
authority's key, it is unlikely that the client would be able to do this
and not make use of the server's certificate key.
The alternative cryptosystem to system used for certificates is the
Diffie--Hellman public--key system\cite{crypto:dh}.
\section{Other sub-protocols}
The protocol described above was the named server protocol, where the
server must provide a certificate. There are two other sub--protocols.
\subsection{Anonymous Server}
In this variant, the server is anonymous and creates a public and private
key pair to be used to establish the session key. It would usually use the
Diffie--Hellman scheme in this case and would simply send to the client the
public key instead of the certificate. This does not weaken the protocol at
all, the client and the server will be able to mutually authenticate one
another, but the server is unknown to the client and to a certification
authority. There is no chain of accountability that could help to locate a
fraudulent server.
\subsection{Named Client and Server}
This variant provides some accountability to the server of the client's
identity and it relies upon the client having a certificate. The protocol
is the same as the named server protocol, but the server can request a
certificate from the client prior to the client sending the pre--master
secret. If the client has no certificate it replies by returning no
certificates, whereupon the server can take its own action, which may well
be to raise an error and not complete the protocol.
\section{Summary}
TLS, like its predecessor the Secure Socket Layer, SSL, does provide both
parties with a mutual belief that the shared session key is a fresh secret.
It also, like SSL, can provide the client with some account of the server's
Internet location and, unlike its predecessor, it does support mutual
authentication certificate exchange. Suffice to say that identities can be
securely established---using X.509 certificates---and that a session key
can be securely established.
A protocol proof is just a basis for a secure implementation. The software
engineering of the authentication protocol has to be considered. An example
of such a failure to ensure that a software implementation was invulnerable
to attack can be found at \cite{sec:pkcs}. The problem with that
implementation was that error messages proved to be too informative
allowing a sophisticated intruder to recover a session key more quickly
than by key trial. Lowe's paper \cite{Lowe96b} has some other
implementation attacks.
\newpage
|
\section{Introduction} \label{sec:Intro}
Some of the most remarkable predictions of quantum field theory arise from
zero point fluctuations of quantum states. One of the best known examples of
this is the Unruh effect \cite{Fulling:73,Davies:75,Unruh:76}, in
which an accelerating detector measures the zero point fluctuations of the
inertial (Minkowski) vacuum state and finds they have a thermal spectrum.
Another is the Casimir effect \cite{Casimir:48}, in which the walls (or
boundaries) of a box experience a net force due to the difference between the
zero point fluctuations of the states inside the box and out.
It seems reasonable, then, to expect accelerating boundaries to produce
interesting effects, and indeed they do. Following studies of the Casimir
effect between moving mirrors in 1+1 dimensions by Moore \cite{Moore:70},
DeWitt pointed out that the single moving mirror problem would be interesting
and could be solved exactly in (1+1)D\cite{DeWitt:74} . Fulling
\cite{Fulling:73} and DeWitt\cite{DeWitt:75} have shown that a {\em uniformly}
accelerating mirror will indeed alter the quantum state in the vicinity of the
mirror. However, a far more interesting result was obtained by Davies and
Fulling\cite{FullingDavies:76,DaviesFulling:77} for a mirror
experiencing {\em non-uniform} acceleration in (1+1)D. Such a mirror actually
emits fluxes of quantum radiation, as though the mirror were knocking zero
point quanta out of the vacuum and off to infinity. More precisely, they found
that a mirror with 2-velocity $u^\mu = d x^\mu/d\tau$ ($u\cdot u = -1$)
and acceleration $a^\mu$ emits a flux
\begin{equation}
\frac{dE}{d\tau} = - \langle T_{\mu\nu}\rangle u^\mu n^\nu =
-\frac{\hbar}{12\pi} \frac{d}{d\tau}~(a\cdot n),
\label{eq:1Dflux}
\end{equation}
in the direction of a unit spatial vector $n^\nu$ orthogonal to $u^\mu$. This
holds for either choice (``left'' or ``right'') of $n^\mu$. Thus, a mirror
whose acceleration is increasing (algebraically) toward the right will emit
a stream of negative energy to the right and a numerically equal positive
stream to the left.
The implications of this result are intriguing. Davies\cite{Davies:72} and
Ford\cite{Ford:78} first raised the possibility that the negative-energy flux
from a moving mirror could be used to cool a hot body and thus violate the
second law of thermodynamics in a quantum context, and this paradox was
further discussed by Deutsch, Ottewill and Sciama\cite{Deutschetal:82}.
Limitations on the extent of such violations in flat spacetime (``quantum
interest'') have been formulated by Ford and Roman\cite{FordRoman:99} and
others (\cite{Flanagan:97}, \cite{Pretorius:99}, {\em etc.}).
More recently, Anderson\cite{Anderson:94} has used this result in the context
of the Geroch gedankenexperiment\cite{Bekenstein:81}. In this experiment, a
box with mirrored walls is filled with radiation and lowered adiabatically
toward a black hole. Unruh and Wald\cite{UnruhWald:82} have shown that such a
box is subject to a buoyancy force and will eventually reach a floating point
above the black hole. Anderson has examined this further and shown that the
ground state inside the box is the Boulware state, whose energy (which becomes
increasingly negative as the box descends) is fed by Davies-Fulling fluxes
from the reflecting walls. This accounts for the buoyancy felt by the box.
The volume of literature on moving mirrors is impressive, but it bears noting
that all the results mentioned above are obtained in 1+1
dimensions. Indeed, if one includes the result for moving mirrors in curved
space-times obtained by Ottewill and Takagi\cite{OttewillTakagi:88} with those
reviewed above, the (1+1)D theory of moving mirrors can be considered
essentially complete. This is due largely to the conformal properties of
quantum field equations in (1+1)D, which allow boundaries, and even
space-time itself, to be flattened, thereby enabling one to obtain results for
complicated geometries from those for much simpler geometries.
This is not the case in 3+1 dimensions, where only partial results are
available. The case of constant acceleration has been solved for both
plane\cite{CandelasDeutsch:77,CandelasDeutsch:78} and
spherical\cite{FrolovSerebriany:79} mirror geometries. Ford and
Vilenkin\cite{FordVilenkin:82} have extended the plane mirror result to
include non-constant acceleration for the case when the acceleration and its
derivatives are small. More recently, Hadasz {\em et al.}\cite{HadaszEtc:98}
have considered arbitrary (radial) motion of a spherical
mirror, but have restricted their attention to the ``S-wave approximation''
where only spherically symmetric modes are considered. Because of this
restriction, their result can be related to the 1+1 dimensional results of
Davies and Fulling\cite{FullingDavies:76,DaviesFulling:77}.
Consideration of quasi-stationary processes (e.g. slow descent of a mirror in
a strong gravitational field) requires knowledge of the flux emitted by a
mirror whose acceleration is changing slowly, though it may be large. Our
objective in this paper is to derive an interesting and relatively simple
result of this type in 3+1 dimensions. The central tool is the use of a
Green's function perturbation technique. Evaluating the perturbation is much
more manageable if the Green's functions for the unperturbed problem are
available in closed form. This is actually the case for a uniformly
accelerated spherical mirror, as shown by Frolov and
Serebriany\cite{FrolovSerebriany:79}. The mirror's history is then a
three-dimensional pseudo-sphere of radius $b$, say, and the unperturbed
problem is just the Minkowski-signature analogue of finding the
four-dimensional electrostatic potential of a point charge in the presence of
an earthed conducting 3-sphere of radius $b$. This is easily solved by the
method of images.
Our objective in this paper is to solve the perturbed Frolov-Serebriany
problem, {\em i.e.} to examine the effect of small spherically symmetric
non-uniformities in the acceleration of a spherical mirror.
It is well to stress at the outset that the solution for a plane mirror cannot
be derived from ours by a straightforward limiting process. The single
parameter $b$, whose reciprocal gives the unperturbed acceleration, also gives
the minimum radius attained by the mirror as seen from its center. Thus, the
plane limit $b\rightarrow\infty$ is inseparable from small acceleration.
(There are reasons to expect the planar case to be considerably more
complicated. Formally, the Wightman Green's function is now an infinite sum of
McDonald (Bessel) functions. Geometrically, any light ray reflected
non-orthogonally of a uniformly accelerated plane mirror will re-encounter the
mirror an infinite number of times; in the spherical case there is just one
encounter.)
Also, we concern ourselves only with calculating the outward flux, which we
expect to be the most interesting stress-energy component. In fact, it turns
out to be somewhat more interesting than one might expect. We find that it
has a remarkable property; although it could, in principle, depend on
the entire retarded history of the mirror, to first order in the mirror
perturbation it depends only the behavior of the mirror at the most recent
retarded time, {\em i.e.} it is local.
This article is organized as follows: in Section \ref{sec:UniAcc} we review
the Frolov-Serebriany result for a mirror expanding with uniform acceleration.
In Section \ref{sec:Perturb} we present our Green's function perturbation
scheme, and in Section \ref{sec:Eval} we use it to evaluate the corrections to
the Frolov-Serebriany Green's function, with some of the more cumbersome
details relegated to Appendix \ref{sec:AppA}. Section \ref{sec:Flux} is
concerned with calculating the quantum flux from these perturbations, with
details again left to Appendices (\ref{sec:AppB} and \ref{sec:AppC}).
Finally, in Section \ref{sec:Conclusion} we offer some concluding remarks.
\section{Uniformly accelerating spherical mirror}\label{sec:UniAcc}
In the case where the mirror's acceleration is uniform, the Green's functions
for the massless fields can be obtained in simple closed form by the method of
images, as noted by Frolov and Serebriany\cite{FrolovSerebriany:79}. In this
Section we shall briefly review these results.
Consider first the static potential due to a point charge $q'$ in Euclidean
4-space at a distance $R'$ from the center of an earthed conducting 3-sphere
of radius $b$. The Dirichlet boundary condition can be reproduced by
introducing a co-radial image charge $q''=-(b/R')^2q'$ at radius $R''=b^2/R'$.
In the Lorentzian analogue of this problem, we are concerned with Green's
functions for the wave equation $\Box \varphi = 0$ in Minkowski space-time,
with Dirichlet boundary conditions on the pseudo-sphere ({\em i.e.} the
time-like hyperboloid of one sheet) $R=b$, where now
\begin{equation}
R^2 \equiv \eta_{\mu\nu}x^\mu x^\nu=x^2+y^2+z^2-t^2=\rho^2+z^2-t^2,
\label{eq:rad}
\end{equation}
in a self-evident notation.
The pseudo-sphere $R=b$ represents the history of a spherical mirror of radius
$b$ (constant as measured in its instantaneous rest frame), whose center is
fixed at the spatial origin $x=y=z=0$ and which moves with uniform
acceleration $a=b^{-1}$.
The image construction gives for the retarded Green's function
$G_{ret}(x,x')$, satisfying
\begin{equation}
\Box G_{ret}(x,x')=-\delta^4(x,x'),
\label{eq:reteq}
\end{equation}
the expression
\begin{equation}
G_{ret}(x,x')=\frac{1}{2\pi}~\theta(t-t')\left\{\delta[(xx')^2]-\left(
\frac{b}{R'}\right)^2\delta[(xx'')^2]\right\}.
\label{eq:Gretconst}
\end{equation}
Here $\theta$ is the unit step (Heaviside) function, $\delta$ is the Dirac
distribution, $(xx')^2$ is the squared Minkowski interval
\begin{equation}
(xx')^2\equiv\eta_{\mu\nu}(x^\mu-{x'}^\mu)(x^\nu-{x'}^\nu),
\label{eq:Minkint}
\end{equation}
and the image source is located at
\begin{equation}
{x''}^\mu = \left(\frac{b}{R'}\right)^2{x'}^\mu.
\label{eq:imagepoint}
\end{equation}
Similarly, the Wightman function
\begin{equation}
W(x,x')=\langle0|\varphi(x)\varphi(x')|0\rangle
\label{eq:Wdef}
\end{equation}
for a massless scalar field takes the form
\begin{equation}
W(x,x')=\frac{1}{4\pi^2}\left\{\frac{1}{(xx')^2+i(t-t')\epsilon} -\left(
\frac{b}{R'}\right)^2\frac{1}{(xx'')^2+i(t-t'')\epsilon}\right\},
\label{eq:Wconst}
\end{equation}
with $\epsilon \rightarrow +0$.
\section{Nearly uniform acceleration: perturbing the boundary}
\label{sec:Perturb}
The corresponding Green's functions for a spherical mirror whose acceleration
is slightly non-uniform can be derived from the preceding results by
superposing the effect of a small perturbation on the history of the mirror,
{\em i.e.}, the time-like 3-space $\Sigma$ on which Dirichlet boundary
conditions are imposed.
Consider generally the problem of solving
\begin{equation}
\Box \Phi = 0, \hspace{1cm} \Phi=0 ~~\mbox{on}~~ \Sigma.
\label{eq:BVP}
\end{equation}
Suppose that $\Sigma$ is a small perturbation of a simpler time-like 3-space
$\Sigma_0$, obtained by displacing $\Sigma_0$ a distance $\delta n(x)$ along
its outward normal $n$, and that we know the solution $\Phi_0$ of the problem
\begin{equation}
\Box \Phi_0 = 0, \hspace{1cm} \Phi=0 ~~\mbox{on}~~ \Sigma_0,
\label{eq:unpertBVP}
\end{equation}
with the same initial boundary conditions.
Then (\ref{eq:BVP}) can be reformulated as a problem with boundary conditions
specified on the unperturbed boundary $\Sigma_0$:
\begin{equation}
\Box \Phi = 0, \hspace{1cm} \Phi=-\frac{\partial\Phi_0}{\partial n}~ \delta n(x)
~~\mbox{ on }~~ \Sigma_0.
\label{eq:pertBVP}
\end{equation}
The causal solution for the perturbation $\delta \Phi \equiv \Phi - \Phi_0$
follows from Green's identity:
\begin{equation}
\delta \Phi(x') = \int_{\Sigma_0} d\Sigma~\frac{\partial G_{ret}(x',x)}{\partial n}
~\delta n(x)~\frac{\partial}{\partial n} \Phi_0(x),
\label{eq:fieldpert}
\end{equation}
where $G_{ret}$ is the retarded Green's function for the unperturbed boundary
value problem (\ref{eq:unpertBVP}) and $x$ is in $\Sigma_0$. Equation
(\ref{eq:fieldpert}) defines a linear operation applied to $\Phi_0$, which we
shall write for brevity as
\begin{equation}
\delta \Phi(x') = L_{\delta n}(x')~\Phi_0(\cdot).
\label{eq:pertop}
\end{equation}
To obtain the effect of the perturbation on the Wightman function
(\ref{eq:Wdef}), we note that it can be written as a mode sum
\begin{equation}
W(x,x') = \int \frac{d^3 k}{(2\pi)^3} ~f_{\boldmath{\mbox{${\scriptstyle k}$}}} (x) ~\overline{f_{\boldmath{\mbox{${\scriptstyle k}$}}}}(x'),
\label{eq:Wsum}
\end{equation}
where $\{ f_{\boldmath{\mbox{${\scriptstyle k}$}}}(x)\}$ is a complete set of initially positive frequency
solutions of (\ref{eq:BVP}) and the bar denotes complex conjugation. Applying
(\ref{eq:pertop}) to each mode separately and summing the results yields
\begin{equation}
\delta W(x,x') = L_{\delta n}(x)W_0(\cdot,x')+L_{\delta n}(x')W_0(x,\cdot),
\label{eq:modepert}
\end{equation}
where $W_0$ denotes the unperturbed Wightman function given by
(\ref{eq:Wconst}).
\section{Wightman function for nearly uniform acceleration}\label{sec:Eval}
Evaluation of the expression (\ref{eq:modepert}) for $\delta W$ in the case
where the mirror's acceleration departs slightly from uniformity is somewhat
lengthy but straightforward. Here, we outline the results of this calculation
reserving the technical details for Appendix \ref{sec:AppA}.
It will be assumed that the mirror remains spherical as viewed by an observer
at its center $x=y=z=0$. Then its history is
\begin{equation}
R=b(1+f(t)), \hspace{1cm} |f(t)| \ll 1
\label{eq:boundpert}
\end{equation}
in terms of this observer's time $t$, where $f(t)$ is arbitrary but small.
The corresponding advanced and retarded times are written
\begin{equation}
v=t+r,\hspace{1cm}u=t-r,\hspace{1cm}r=\sqrt{x^2+y^2+z^2}.
\label{eq:uvdef}
\end{equation}
The radial energy flux measured by a stationary observer outside the mirror is
\begin{equation}
F= T_{uu}-T_{vv}.
\label{eq:statflux}
\end{equation}
Each of these terms can be dealt with by a similar procedure. Let us consider
$T_{uu}$; it is evident from (\ref{eq:Wdef}) that its expectation value at an
event $p_1$ outside the mirror is
\begin{equation}
\langle T_{uu}(p_1)\rangle = \left[ \partial_{u_1} \partial_{u_2} W(p_1,p_2)
\right]_{p_2=p_1}
\label{eq:fluxdef}
\end{equation}
for a minimally coupled massless scalar field, and a similar but more
complicated expression for conformal coupling (see (\ref{eq:quantSETconf})
below). Note that regularization and symmetrization are not needed to evaluate
this component of the stress-energy.
We can now present the results for $\delta W$. The spherical symmetry of
(\ref{eq:boundpert}) allows us to take both $p_1$ and $p_2$ in the $z-t$ plane
($x=y=0$). Calculation shows (see Appendix \ref{sec:AppA}) that the first
term of (\ref{eq:modepert}) contributes
\begin{equation}
\frac{8\pi^2\zeta^2}{(R_1^2-b^2)(R_2^2-b^2)} L_{\delta n}(p_1)W(\cdot,p_2)
= \frac{z_1z_2}{\zeta} \int_{-\infty}^{t_1^*} dt \frac{f(t)}{(t-t_0)^3}
-\frac{z_1z_1^*}{(R_1^2-b^2)(t_1^*-t_0)^2}
\label{eq:Wterm1}
\end{equation}
where $\zeta$ and $t_0$ are defined by
\begin{equation}
\zeta=z_1t_2-z_2t_1, ~~~~~\zeta t_0 = \frac{1}{2}(z_1-z_2)(z_1z_2-b^2)
+(z_1t_2^2-z_2t_1^2),
\label{eq:zetadef}
\end{equation}
and $p_i^*$ (with coordinates $x_i^*,t_i^*$) is the event nearest to $p_i$
($i=1,2$) at which the past light-cone of $p_i$ intersects the (unperturbed)
mirror (see Fig. 1).
\begin{figure}[hbtp]
\begin{center}
\leavevmode \epsfbox{2d_intersect.eps}
\end{center}
\caption{The $z-t$ plane. The mirror profile is the hyperbola. $p_1$ and $p_2$
are the two events at which the Wightman function is to be evaluated and $v_0$
is their common advanced time. $p_1^*$ and $p_2^*$ are the intersections of
their past light-cones with the mirror in the $z-t$ plane.
\label{fig:1}}
\end{figure}
The contribution of the second term in (\ref{eq:modepert}) is obtained by
interchanging $p_1$ and $p_2$ in (\ref{eq:Wterm1}). Because $\zeta$ is an odd
function of $p_1,~p_2$ ($t_0$ is even), the term involving the integral
changes sign. Thus, the sum of the two contributions,
\begin{equation}
-\frac{8\pi^2\zeta^2}{(R_1^2-b^2)(R_2^2-b^2)}\delta W(p_1,p_2) =
\frac{z_1z_2}{\zeta} \int_{t_1^*}^{t_2^*} dt \frac{f(t)}{(t-t_0)^3}
+\sum_{i=1}^2\frac{z_iz_i^*}{(R_i^2-b^2)(t_i^*-t_0)^2}f(t_i^*),
\label{eq:deltaW}
\end{equation}
depends only on the mirror's history between the retarded times $t_1^*$ and
$t_2^*$. Particle and anti-particle modes interfere destructively in the case
of a spherical mirror to eliminate the effects of the past history and
produce (when we take the coincidence limit $p_2 \rightarrow p_1$) a
purely local expression for the stress-energy.
To perform the partial derivatives $\partial_{u_1}$ and $\partial_{u_2}$ in the
expression (\ref{eq:fluxdef}) for~~$T_{uu}$, we require mutual independence of
the $u$-coordinates of the events $p_1$ and $p_2$. But these are the only
coordinates which need be independent. To evaluate $T_{uu}$, it suffices to
consider $\delta W(p_1,p_2)$ in a partial pre-coincidence limit $v_1=v_2$, as
in Fig. \ref{fig:1}.
These remarks in principle apply, {\em mutatis mutandis}, also to the
evaluation of~~$T_{vv}$, but there is a complication. In the pre-coincidence
limit $u_2\rightarrow u_1$, the points $p_1^*$ and $p_2^*$ tend to coincidence
with each other and with the point on the mirror having coordinate $t_0$, so
that the integrand in (\ref{eq:deltaW}) becomes infinite while the interval of
integration shrinks to zero. The evaluation of $T_{vv}$ is discussed further in
Appendix \ref{sec:AppC}. Here, we merely note that $T_{vv}$ makes no
radiative contribution (proportional to $r^{-2}$) to the flux
(\ref{eq:fluxdef}). This follows at once from the identity
\begin{equation}
\partial_u\left(r^2 \partial_u \left(r^2T_{vv}\right)\right) = \partial_v
\left( r^2 \partial_v \left( r^2 T_{uu}\right)\right),
\label{eq:consEq}
\end{equation}
which is a consequence of the conservation of $T_{\mu \nu}$ and the vanishing
of the trace $T^\alpha_\alpha$ for a conformal scalar field in flat space. In
(\ref{eq:consEq}), the derivative $\partial_v$ increases the fall-off with
distance, but this does not hold for $\partial_u$, which can operate on the
retarded displacement $f(u)$ in (\ref{eq:boundpert}). Thus, $T_{vv}$ falls off
more strongly than $T_{uu}$. The detailed calculation (Appendix
\ref{sec:AppC}) shows that $T_{vv}\sim r^{-6}$ as $r\rightarrow\infty$.
\section{Flux from a non-uniformly accelerating mirror}\label{sec:Flux}
The expectation value of the stress-energy is derivable from the partial
derivatives of the Wightman function $W(x,x')=W_0+\delta W$ in the coincidence
limit, with the unperturbed part $W_0$ given by (\ref{eq:Wconst}) and the
perturbation $\delta W$ by (\ref{eq:deltaW}). The unperturbed part becomes
singular in the limit $x' \rightarrow x$, but is easily regularized by
subtracting the value of $W_0$ in free space without the mirror, {\em i.e.},
the first term of (\ref{eq:Wconst}), leaving the second (image) term as the
sole contribution to $(W_0)_{reg}$. The perturbation $\delta W$ is regular in
the coincidence limit.
For a massless scalar field, two different stress tensors are commonly
considered: (a) the minimal stress-energy, given classically by
\begin{equation}
\left(T_{\mu \nu}\right)_{min} = \varphi_{,\mu} \varphi_{,\nu} -
\frac{1}{2} g_{\mu \nu} (\nabla \varphi)^2
\label{eq:classSETmin}
\end{equation}
and quantum mechanically by
\begin{equation}
\langle T_{\mu \nu}(x)\rangle_{min} = \left[(\partial_{\mu}\partial_{\nu'}-\frac{1}{2}
g_{\mu\nu}~\partial^{\alpha}\partial_{\alpha'})W_{reg}(x,x')\right]_{sym~;~x'=x},
\label{eq:quantSETmin}
\end{equation}
in which ``sym'' indicates symmetrization in (x,x') and in the partial
derivatives; (b) the conformal (trace-free) stress-energy, defined classically
by
\begin{equation}
\left(T_{\mu \nu}\right)_{\mathit conf} = \frac{2}{3}\varphi_{,\mu}
\varphi_{,\nu}- \frac{1}{3}\varphi\varphi_{,\mu\nu}-
\frac{1}{6}g_{\mu \nu} (\nabla \varphi)^2
\label{eq:classSETconf}
\end{equation}
and quantum mechanically by
\begin{equation}
\langle T_{\mu \nu}(x)\rangle_{\mathit conf}=\frac{1}{3}\left[
(2\partial_{\mu}\partial_{\nu'}- -\partial_{\mu'}\partial_{\nu'}
-\frac{1}{2} g_{\mu\nu}~\partial^{\alpha}\partial_{\alpha'})
W_{reg}(x,x')\right]_{sym~;~x'=x},
\label{eq:quantSETconf}
\end{equation}
We begin by reviewing the Frolov-Serebriany\cite{FrolovSerebriany:79} results
for {\em uniform} acceleration. Differentiating the regularized form of
(\ref{eq:Wconst}), we easily find
\begin{eqnarray}
\langle T^{(0)}_{\mu \nu}(x)\rangle_{min} &=& - \frac{b^2}{\pi^2(R^2-b^2)^4}
\left(x^{\mu}x^{\nu}-\frac{1}{2}g^{\mu\nu}R^2\right),
\label{eq:constSETmin}\\
\langle T^{(0)}_{\mu \nu}(x)\rangle_{\mathit conf} &=& 0.
\label{eq:constSETconf}
\end{eqnarray}
This last result is quite remarkable, because the conformal stress is not
likely to vanish for a spherical mirror at rest (it certainly does not for
electromagnetic fields \cite{Boyer:68,Davies:72,MiltonEtc:78}).
It appears that the effects of uniform acceleration exactly cancel the static
Casimir stresses.
The effects of non-uniform acceleration are more complicated. We shall simply
quote the result for the conformal radial out-flux $\langle T_{uu}
\rangle_{\mathit conf}$ at a point $(r,t)$ outside the mirror, leaving to
Appendix \ref{sec:AppB} an outline of the derivation:
\begin{equation}
\langle T_{uu}(t,r) \rangle_{\mathit conf} = - \frac{1}{1440 \pi^2}
\frac{1}{rv} \left\{\frac{q}{m^3n}\frac{d^2\alpha}{d\chi^2} -
\frac{4}{m^3n^2r} (npr-q^2)\frac{d \alpha}{d \chi} +
\frac{2 s^2}{mn^3r^2}\left[q(\alpha+f)+p\frac{df}{d\chi}\right]\right\}.
\label{eq:SETpert}
\end{equation}
The notation is as follows: the advanced and retarded times, $v$ and $u$ are
defined as in (\ref{eq:uvdef}), and we write
\begin{eqnarray}
m&=&-\frac{u}{2},~~~n~=~\frac{1}{2v}(R^2-b^2)~=~-\frac{1}{2}(u+b^2/v),
\nonumber\\
p&=&\frac{1}{8}(u^2-b^2),~~~q~=~\frac{1}{8}(u^2+b^2),~~~
s~=~-\frac{1}{2}(v+b^2/v);
\label{eq:notdefs}
\end{eqnarray}
$\chi$ is the pseudo-angle along the (unperturbed) mirror trajectory ({\em
i.e.}, $\tau=b\chi$ is the mirror's proper time). Equation
(\ref{eq:boundpert}), giving the trajectory of the perturbed mirror, is now
written $R=b(1+f(\chi))$, and
\begin{equation}
\alpha = f''(\chi) - f(\chi)
\label{eq:alphadef}
\end{equation}
is a measure of the non-uniformity of the acceleration $A$, which is given by
\begin{equation}
A = (1+\alpha)/b.
\label{eq:acceldef}
\end{equation}
In (\ref{eq:SETpert}), $\chi$ refers to the pseudo-angle at the point $p^*$,
which is the nearest retarded point to (r,t), as in Fig. \ref{fig:2}.
The corresponding expression for the canonical flux is too long to reproduce
here, and is also deferred to Appendix B.
\begin{figure}[hbtp]
\begin{center}
\leavevmode \epsfbox{3d_intersect.eps}
\end{center}
\caption{Intersection of the history of a uniformly accelerating sphere
(hyperboloidal cylinder) and the past null cone of an exterior point $p$. The
intersection (which lies entirely within the shaded plane) is represented by
the bold curve. The nearest retarded point to $p$ on the mirror's worldsheet,
$p^*$, is the only point whose perturbation contributes to the flux at $p$.
This figure has been dimensionally reduced; each point represents a circle.
\label{fig:2}}
\end{figure}
The limit $b \rightarrow \infty$ corresponds to a slowly accelerating, nearly
plane mirror. This was the case studied by Ford and
Vilenkin\cite{FordVilenkin:82}. It is straightforward to show that in this
limit our result (\ref{eq:SETpert}) for the conformal flux (and also our
result for the canonical flux) reduces to the expressions they give.
To obtain a more intuitive grasp of the physical meaning of the complex
expression (\ref{eq:SETpert}), we can evaluate the flux $F$ radiated at
retarded time $u=-b$ ~~-~~ {\em i.e.} when the mirror is near its minimum
radius $r_M \approx b$ ~~-~~ as measured by a stationary observer at radius
$r\gg b$ and the same retarded time. Using (\ref{eq:statflux}), taking the
appropriate limit of (\ref{eq:SETpert}), and noting that $T_{vv}$ does not
contribute to the flux to leading order (as discussed at the end of Section
\ref{sec:Eval}) we find
\begin{equation}
F=-\frac{\hbar}{720 \pi^2}\left(\frac{R_0}{r}\right)^2\left\{A
\frac{d^2 A}{d \tau^2}+2A^3\left(A-\frac{1}{R_0}\right)\right\},
\label{eq:flux}
\end{equation}
where $R_0=b(1+f(0))$ is the proper radius of the mirror at the time of
emission $\chi = 0$, and we have restored Planck's constant to display the
correct dimensionality. We recall that this perturbative result is correct
to linear order in deviations from uniform acceleration $(A-1/R)$, $dA/d\tau$,
$d^2A/d\tau^2$, but the acceleration $A$ itself is arbitrary.
\section{Concluding Remarks}\label{sec:Conclusion}
Our chief interest in this paper has been in the quantum flux radiated by the
mirror and we have explicitly computed only those components of the
stress-energy tensor from which it arises. However, the remaining components
can be derived straightforwardly (though with some labour) from our expression
(\ref{eq:deltaW}) with the methods of Appendix \ref{sec:AppB}. It would be
useful to have these to round out the picture.
The relevant Green's functions for an unperturbed, uniformly accelerating
spherical mirror have a simple closed form, and this has enormously simplified
our perturbative calculation. This simplification is bought at a price: we are
limited to spherical mirrors whose acceleration $A$ and proper radius $R$ are
nearly reciprocals. We cannot decouple the plane limit $R\rightarrow \infty$
from the limit of small acceleration, and cannot disentangle curvature
(Casimir) effects from the effects of acceleration.
It is evident that much remains to be done before we can claim anything
approaching a comprehensive understanding of the quantum dynamics of
three-dimensional mirrors.
\acknowledgements{We are indebted to Valeri Frolov and Tom Roman for
discussions and to the latter for calling our attention to the prior work of
Ford and Vilenkin \cite{FordVilenkin:82} on slowly accelerating plane mirrors.
This research was supported by the Canadian Institute for Advanced Research
and by NSERC of Canada.}
|
\section{Introduction}
It is well known for a long time that a point charge at rest in a static spacetime
feels an electrostatic self-force. The calculation is performed by considering
the global electrostatic potential determined as the solution of the
Maxwell equations in the background metric of the spacetime. However, it would
seem that its existence was a curiosity. The situation has recently undergone a
change when Bekenstein and Mayo \cite{bek1}
and Hod \cite{hod} have derived the upper entropy bound for a charged object by
requiring the validity of thermodynamics of the Reissner-Nordstr\"{o}m
black hole. Their proof takes really into account the expression of the
electrostatic self-energy for a point charge at rest in a Schwarzschild
black hole which has been previously determined in closed form
\cite{smi1,zel,lea1}.
The purpose of this work is to extend these results to a new case where it is possible
to determine explicitly the electrostatic self-energy. We consider the
spacetime, introduced by Aryal {\em et al} \cite{ary}, which
describes a Schwarzschild black hole pierced by a cosmic string.
It represents a straight cosmic string, infinitely thin, passing through a
spherically symmetric black hole. It is
obtained by cutting a wedge in the Schwarzschild geometry. So,
in the coordinate system $(t,r,\theta ,\varphi )$ with $0\leq \varphi <2\pi$,
the metric can be written
\begin{equation}\label{Sch}
ds^2=-\left( 1-\frac{2m}{r}\right) dt^2+\left( 1-\frac{2m}{r}\right)^{-1}dr^2
+r^2d\theta^2+B^2r^2\sin^2\theta d\varphi^2
\end{equation}
where $m$ is a positive parameter and $B$ is related to the linear mass density
$\mu$ of the cosmic string by $B=1-4\mu$ with $0<B<1$. We only consider
the spacetime outside the horizon, i.e. for $r>2m$.
In section 2, we summarise the Maxwell equations in metric (\ref{Sch})
and, in the case $1/2<B<1$, we give explicitly the expression of the electrostatic
potential generated by a point charge at rest. The
proof that this expression obeys the electrostatic equation is fulfilled in section 3.
Taking into account the found expression of the electrostatic self-energy,
we derive in section 4 the entropy bound for a charged object by employing
thermodynamics of the black hole. We add some concluding remarks in
section 5.
\section{Electrostatic potential}
The Maxwell equations in metric (\ref{Sch}), having as source a point
charge $e$ located at the position $(r_0,\theta_0,\varphi_0)$ with $r_0>2m$,
reduce to
\begin{equation}\label{Max}
\partial_i\left( \sqrt{-g}F^{i0}\right) =-\frac{e}{4\pi}
\delta (r-r_0)\delta (\theta -
\theta_0)\delta (\varphi -\varphi_0) \quad {\rm with}\quad
F_{i0}=\partial_iA_0
\end{equation}
where $A_0$ is the electrostatic potential.
According to (\ref{Max}), the electrostatic equation for $A_0$ can be written
\begin{eqnarray}\label{elect}
\nonumber & & \frac{1}{r^2}\frac{\partial}{\partial r}\left( r^2\frac{\partial}{\partial r}
A_0\right) +\frac{1}{r(r-2m)\sin \theta}\frac{\partial}{\partial \theta}
\left( \sin \theta \frac{\partial}{\partial \theta}A_0\right)
+\frac{1}{B^2r(r-2m)}\frac{\partial^2}{\partial \varphi^2}A_0 \\
& & =-\frac{e}{4\pi Br^2\sin \theta}\delta (r-r_0)\delta (\theta -\theta_0)
\delta (\varphi -\varphi_0) \, .
\end{eqnarray}
We point out that the application of the Gauss theorem to
equation (\ref{Max}) yields
\begin{equation}\label{gauss}
A_0(r,\theta ,\varphi )\sim \frac{e}{Br} \quad {\rm as}\quad
r\rightarrow \infty
\end{equation}
if there is no electric flux through the horizon.
Furthermore, we require that the electromagnetic field is regular at the
horizon by imposing that $F^{\mu \nu}F_{\mu \nu}$ is finite as $r\rightarrow 2m$.
We limit ourselves to the case where $1/2<B<1$ which is physically justified
for a cosmic string since $\mu \ll 1$. Without loss of generality,
we put $\varphi_0=\pi$ to simplify. The expression of the electrostatic potential $A_0$
satisfying equation (\ref{elect}) with the desired boundary conditions can
be expressed as the following sum
\begin{equation}\label{ssc}
A_0(r,\theta ,\varphi )=V^*(r,\theta ,\varphi )+
V_B(r,\theta ,\varphi )+\frac{em}{Brr_0}
\end{equation}
where the expressions of $V^*$ and $V_B$ are given below.
\begin{figure}
\begin{picture}(320,100)(10,10)
\put(10,58){$\varphi =\pi$}
\put(50,60){\line(1,0){10}}\put(70,60){\line(1,0){10}}\put(90,60){\line(1,0){10}}
\put(110,60){\line(1,0){10}}\put(130,60){\line(1,0){10}}\put(150,60){\line(1,0){10}}
\put(170,60){\line(3,1){120}}
\put(300,100){$\varphi =\pi /B-\pi$}
\put(300,12){$\varphi =3\pi -\pi /B$}
\put(310,65){$\varphi =0$}\put(310,53){$\varphi =2\pi$}
\put(170,60){\line(3,-1){120}}
\put(170,60){\line(10,0){130}}
\put(237,70){\underline{$n=1$}}\put(235,48){\underline{$n=-1$}}
\put(110,75){\underline{$n=0$}}
\end{picture}
\caption{Regions delimited by the hypersurfaces $\varphi =$ constant}
\end{figure}
To express the potential $V^*$, we must consider the regions of the spacetime
delimited by the hypersurfaces $\varphi =$ constant as shown on Figure~1.
We have
\begin{equation}\label{vetoile}
V^*(r,\theta ,\varphi )=\left\{ \begin{array}{ll}
V_C(r,\sigma_0(\theta ,\varphi )]+V_C[r,\sigma_1(\theta ,\varphi )]
& \quad 0<\varphi <\pi /B -\pi \\
V_C[r,\sigma_0(\theta ,\varphi )] & \quad \pi /B-\pi <\varphi <3\pi -\pi /B \\
V_C[r,\sigma_0(\theta ,\varphi )]+V_C[r,\sigma_{-1}(\theta ,\varphi )]
& \quad 3\pi -\pi /B<\varphi <2\pi
\end{array} \right.
\end{equation}
where $V_C$ is the Copson potential \cite{cop} which is a solution to
electrostatic equation (\ref{elect}) with $B=1$, i.e. for the
Schwarzschild black hole. Its expression is
\begin{equation}\label{copson}
V_C[r,\sigma ]=\frac{e}{rr_0}\frac{(r-m)(r_0-m)-m^2\sigma}
{[(r-m)^2+(r_0-m)^2-m^2-2(r-m)(r_0-m)\sigma +m^2\sigma^2]^{1/2}}
\end{equation}
The variables $\sigma_n$ in formula (\ref{vetoile})
are the following functions of $\theta$ and $\varphi$
\begin{eqnarray}\label{sig0}
\nonumber & & \sigma_0(\theta ,\varphi )=\cos \theta \cos \theta_0+\sin \theta
\sin \theta_0\cos B(\varphi -\pi ) \, , \\
& & \sigma_1(\theta ,\varphi )= \cos \theta \cos \theta_0+\sin \theta
\sin \theta_0\cos B(\varphi +\pi ) \, , \\
\nonumber & & \sigma_{-1}(\theta ,\varphi )=\cos \theta \cos \theta_0+\sin \theta
\sin \theta_0 \cos B(\varphi -3\pi ) \, .
\end{eqnarray}
The potential $V_B$ is given by the integral expression
\begin{equation}\label{vb}
V_B(r,\theta ,\varphi )=\frac{1}{2\pi B}\int_{0}^{\infty}
V_C[r,k(\theta ,x)]F_B(\varphi ,x)dx
\end{equation}
where the function $k$ is given by
\begin{equation}\label{kx}
k(\theta ,x)=\cos \theta \cos \theta_0-\sin \theta \sin \theta_0\cosh x
\end{equation}
and the function $F_B$ by
\begin{equation}\label{fb}
F_B(\varphi ,x)=-\frac{\sin (\varphi -\pi /B)}
{\cosh x/B+\cos (\varphi -\pi /B)}
+\frac{\sin (\varphi +\pi /B)}{\cosh x/B+\cos (\varphi +\pi /B)} \, .
\end{equation}
In the Schwarzschild black hole, sum (\ref{ssc}) with
$V_1=0$ and $V^*=V_C$ yields the electrostatic potential that we have
already obtained \cite{lin1}. On the other hand for the
cosmic string, i.e. $m=0$, we find our previous result \cite{lin2}, already
known in the case of a wedge in flat space \cite{gar,obe}.
The electrostatic self-potential in a neighbourhood of the point charge is
$A_0(r,\theta ,\varphi )-V_C[r,\sigma_0(\theta ,\varphi )]$. In
consequence, the electrostatic self-energy $W_{self}$ is
\begin{equation}\label{self}
W_{self}(r_0,\theta_0)=\frac{e^2m}{2Br_{0}^{2}}
-\frac{e\sin \pi/B}{2\pi B}\int_{0}^{\infty}
V_C[r_0,k(\theta_0,x)]\frac{dx}{\cosh x/B-\cos \pi /B} \, .
\end{equation}
From (\ref{self}), we can deduce the electrostatique self-force which has
been already obtained in the Schwarzschild black hole \cite{smi1,zel,lea1}
and in the cosmic string \cite{lin3,smi2}.
\section{Checking of the electrostatic solution}
We must firstly verify that sum (\ref{ssc}) is a solution to
equation (\ref{elect}). The potential $V^*$ is obviously a local solution
since the Copson potential $V_C$ expressed in variables $r$, $\theta$
and $\phi$ with $\phi =B\varphi$ obeys the electrostatic equation for the
Schwarzschild black hole. As a consequence, the function
$V_C[r,k(\theta ,x)]$ satisfies
\[
\frac{1}{r^2}\frac{\partial}{\partial r}\left( r^2\frac{\partial}{\partial r}
V_C\right)+\frac{1}{r(r-2m)\sin \theta}\frac{\partial}{\partial \theta}
\left( \sin \theta \frac{\partial}{\partial \theta}V_C\right)
-\frac{1}{r(r-2m)}\frac{\partial^2}{\partial x^2}V_C=0 \, .
\]
Then, according to its expression (\ref{vb}), potential $V_B$ obeys
electrostatic equation (\ref{elect}) without second member if we have
\[
\int_{0}^{\infty}\frac{\partial^2}{\partial x^2}V_C[r,k(\theta ,x)]
F_B(\varphi ,x)dx+\frac{1}{B^2}\int_{0}^{\infty}V_C[r,k(\theta ,x)]
\frac{\partial^2}{\partial \varphi^2}F_B(\varphi ,x)dx =0 \, .
\]
Now from expression (\ref{fb}), we verify that
\begin{equation}\label{resc}
\frac{\partial^2}{\partial x^2}F_B(\varphi ,x)+\frac{1}{B^2}
\frac{\partial^2}{\partial \varphi^2}F_B(\varphi ,x)=0
\end{equation}
which ensures the condition mentioned above after two
sucessive integrations by part.
We notice that the potential $em/Brr_0$ is a homogeneous solution
to electrostatic equation which is regular at the horizon.
We secondly check that sum (\ref{ssc}) is continuous. At $\varphi =0$,
it is clear because $V^*(r,\theta ,0)=V^*(r,\theta ,2\pi )$ since
$\sigma_1(r,\theta ,0)=\sigma_{-1}(r,\theta ,2\pi )$.
At $\varphi =\pi /B-\pi$, we introduce $\epsilon$ by setting
$\varphi =\pi /B-\pi +\epsilon$ and then the potential $V_B$ becomes
\begin{eqnarray}\label{epsilon}
\nonumber & & V_B(r,\theta ,\pi /B-\pi +\epsilon )=
\frac{\sin \epsilon}{2\pi B}\int_{0}^{\infty}
V_C[r,k(\theta,x)]\frac{dx}{\cosh x/B -\cos \epsilon} \\
& & -\frac{\sin (2\pi /B+\epsilon )}{2\pi B}\int_{0}^{\infty}V_C[r,k(\theta ,x)]
\frac{dx}{\cosh x/B-\cos (2\pi /B+\epsilon)} \, .
\end{eqnarray}
We write down the following integral
\[
\sin \epsilon \int_{0}^{x}\frac{dy}{\cosh y-\cos \epsilon}=
2\arctan \left( \tanh \frac{x}{2}\cot \frac{\epsilon}{2}\right)
\quad {\rm with}\quad \epsilon \neq 0 \, .
\]
By integrating by part the first term of expression (\ref{epsilon}), we get
\[
\frac{1}{2}V_C[r,k(\theta ,\infty )]-\frac{1}{\pi}\int_{0}^{\infty}
\frac{\partial}{\partial y}V_C[r,k(\theta ,By)]\arctan \left(
\tanh \frac{y}{2}\cot \frac{\epsilon}{2}\right) dy \, .
\]
But the function $\arctan$ is bounded by $\pi /2$ and consequently we may
take the limit $\epsilon \rightarrow 0$ inside the integral. We obtain thereby
\[
\frac{1}{2}V_C[r,k(\theta ,\infty )]-\frac{1}{2}\left\{
V_C[r,k(\theta ,\infty )]-V_C[r,k(\theta ,0)]\right\}
\quad {\rm as}\quad \epsilon \rightarrow 0 \; \epsilon >0 \, .
\]
We have thus prove that integral expression (\ref{epsilon}) verifies
\begin{eqnarray}\label{limite}
\nonumber & & \lim_{\epsilon \rightarrow 0\; \epsilon >0}V_B
(r,\theta ,\pi /B-\pi +\epsilon )=\frac{1}{2}V_C[r,k(\theta ,0)] \\
& & -\frac{\sin 2\pi /B}{2\pi B}\int_{0}^{\infty}
V_C[r,k(\theta ,By)]\frac{dy}{\cosh y-\cos 2\pi /B}
\end{eqnarray}
and another with $\epsilon <0$ yielding $-V_C/2$ in formula (\ref{limite}).
On the other hand, the potential $V^*$ verifies
\begin{equation}\label{disc}
\lim_{\epsilon \rightarrow 0 \; \epsilon >0} \left(
V^*(r,\theta ,\pi /B-\pi +\epsilon )-V^*(r,\theta ,\pi /B-\pi -\epsilon )
\right) = -V_C[r,k(\theta ,0)] \, .
\end{equation}
By combining results (\ref{limite}) and (\ref{disc}), we thus obtain that
the potentials $V_B$ and $V^*$ are both discontinuous at
$\varphi =\pi /B -\pi$ whereas their sum is regular. Of course,
the potential $V^*+V_B$ is also continuous at $\varphi =3\pi -\pi /B$ by symmetry.
In conclusion, the electrostatic potential $A_0$ is a smooth function only
singular at the position of the point charge.
Furthermore, it is easy to show
that the derivative of the electrostatic potential $A_0$ with
respect to $\varphi$ is everywhere continuous.
We thirdly determine the asymptotic form of the electrostatic potential $A_0$.
From expression (\ref{vetoile}) of $V^*$, we have immediately
\begin{equation}\label{v*asymp}
V^*(r,\theta ,\varphi )\sim \left\{ \begin{array}{ll}
2e(r_0-m)/rr_0 & \quad 0<\varphi <\pi /B -\pi \\
e(r_0-m)/rr_0 & \quad \pi /B -\pi <\varphi <3\pi -\pi /B \quad
{\rm as}\quad r\rightarrow \infty \\
2e(r_0-m)/rr_0 & \quad 3\pi -\pi /B <\varphi <2\pi \end{array} \right. \, .
\end{equation}
On the other hand, from expression (\ref{vb}) of $V_B$ we get
\begin{equation}\label{vbasymp}
V_B(r,\theta ,\varphi )\sim
\frac{e(r_0-m)}{rr_0}g(\varphi ) \quad {\rm as}\quad r\rightarrow \infty
\end{equation}
with
\[
g(\varphi )=\frac{1}{2\pi}\int_{0}^{\infty}\left[
\frac{\sin (\varphi +\pi /B)}{\cosh x+\cos (\varphi +\pi /B)}
-\frac{\sin (\varphi -\pi /B)}{\cosh x+\cos (\varphi -\pi /B)}\right] dx
\]
which can be integrated by elementary methods
\begin{equation}\label{gdyexp}
g(\varphi )= \left\{ \begin{array}{ll}
1/B-2 & \quad 0<\varphi < \pi /B-\pi \\
1/B-1 & \quad \pi /B-\pi <\varphi < 3\pi -\pi /B \\
1/B-2 & \quad 3\pi -\pi /B <\varphi <2\pi \end{array} \right. \, .
\end{equation}
By using (\ref{v*asymp}) and (\ref{vbasymp}) with (\ref{gdyexp}), we obtain that
the electrostatic potential (\ref{ssc}) has the desired asymptotic form (\ref{gauss}).
At last, we must verify that the electromagnetic field derived from the
electrostatic potential (\ref{ssc}) is regular at the horizon.
This point results of the fact that $V_C$ tends to $e/rr_0$ when
$r\rightarrow 2m$.
\section{Entropy bound for a charged object}
We now consider the spacetime which describes a Reissner-Nordstr\"{o}m
black hole pierced by a cosmic string. It is obtained by cutting a wedge
in the Reissner-Nordstr\"{o}m geometry. In the coordinate
$(t,r,\theta ,\varphi )$ with $0\leq \varphi <2\pi$, the metric can be written
\begin{eqnarray}\label{RN}
\nonumber & & ds^2=-\left( 1-\frac{2E}{Br}+\frac{q^2}{B^2r^2}\right) dt^2 \\
& & +\left( 1-\frac{2E}{Br}+\frac{q^2}{B^2r^2}\right)^{-1}dr^2+
r^2d\theta^2+B^2r^2\sin^2\theta d\varphi^2
\end{eqnarray}
where $E$ and $q$ are two parameters. We only consider the spacetime
outside the outer horizon, i.e. $r>(E+\sqrt{E^2-q^2})/B$ by assuming that
$E^2>q^2$. Following \cite{ary,mar}, we interpret $E$ as the energy of
the black hole. Clearly, $q$ is the electric charge of the black hole.
For $q=0$, metric (\ref{RN}) reduces to metric (\ref{Sch})
by setting $m=E/B$. The horizon area ${\cal A}$ of the black hole defined by
metric (\ref{RN}) has the expression
\begin{equation}
{\cal A}(E,q)=\frac{4\pi}{B}\left( E+\sqrt{E^2-q^2}\right)^2
\end{equation}
and the entropy $S_{BH}$ of the black hole is given by
\begin{equation}\label{aire}
S_{BH}(E,q)=\frac{1}{4}{\cal A}(E,q) \, .
\end{equation}
The Reissner-Nordstr\"{o}m black hole pierced by a cosmic string linearised
with respect to its electric charge $q$ is described by metric (\ref{Sch})
plus an electromagnetic test field having the electrostatic potential
\begin{equation}\label{ext}
A_{0}^{ext}(r,\theta ,\varphi )=\frac{q}{Br} \, .
\end{equation}
Moreover, the black hole entropy (\ref{aire}) reduces to
\begin{equation}\label{entropy}
S_{BH}(E,q) \approx \frac{2\pi}{B} \left( 2E^2-q^2\right) \, .
\end{equation}
The original method of Bekenstein \cite{bek2} for finding the entropy bound
for a neutral object in the Schwarzschild black hole has been recently
extented for charged object in the
Reissner-Nordstr\"{o}m black hole \cite{bek1,hod}. Referring to
\cite{bek1,hod}, we
recall that the energy ${\cal E}$ of a charged object with a mass $\mu$, an
electric charge $e$ and a radius $R$ located at the position
$(r_0,\theta_0)$ in metric (\ref{Sch}), in presence of the exterior
electrostatic potential (\ref{ext}), has the expression
\begin{equation}\label{energy}
{\cal E}=\sqrt{1-\frac{2E}{Br_0}}+\frac{eq}{Br_0}+
W_{self}(r_0,\theta_0)
\end{equation}
where $W_{self}$ is the electrostatic self-energy (\ref{self}).
When the charged object is just outside the horizon,
its energy (\ref{energy}), for a very small proper length $R$, is
\begin{equation}\label{last}
{\cal E}_{last}\sim \frac{\mu RB}{4E}+\frac{eq}{2E}+W_{self}(2E/B,\theta_0)
\quad {\rm as}\quad R\rightarrow 0 \, .
\end{equation}
In this state, the system formed by the black hole and the charged object has
an entropy $S_{BH}(E,q)+S$ where $S$ is
the entropy of the charged object. When the charged object falls in the
horizon, the final state is a Reissner-Nordstr\"{o}m black hole with the new parameters
\begin{equation}
E_f=E+{\cal E}_{last} \quad {\rm and}\quad q_f=q+e \, .
\end{equation}
But in this final state, the entropy is $S_{BH}(E_f,q_f)$. We now write down
the generalised second law of thermodynamics
\begin{equation}\label{ineg}
S_{BH}(E_f,q_f)\geq S_{BH}(E,q)+S \, .
\end{equation}
We can calculate $\triangle S_{BH}=S_{BH}(E_f,q_f)-S_{BH}(E,q)$ from
expression (\ref{entropy}). We keep only linear terms in ${\cal E}_{last}$.
By this way, we thus exclude a possible
gravitational self-force which should be quadratic in $\mu$
as in a cosmic string \cite{smi2}. We find
\begin{equation}\label{triangle}
\triangle S_{BH}=\frac{4\pi}{B} \left[2E{\cal E}_{last}-eq-
\frac{e^2}{2} \right] \, .
\end{equation}
By inserting (\ref{last}) into (\ref{triangle}), we get
\begin{equation}\label{bound}
\triangle S_{BH}=\frac{4\pi}{B}\left[ \frac{\mu RB}{2}+2E\left(
\frac{e^2}{8E}+\frac{e^2B}{8E}g(\pi ) \right)
-\frac{e^2}{2} \right] \, .
\end{equation}
where $g(\pi )=1/B-1$ by formula (\ref{gdyexp}).
According to inequality (\ref{ineg}), we obtain then from (\ref{bound})
the desired entropy bound
\begin{equation}
S \leq 2\pi \left[ \mu R-\frac{e^2}{2}\right] \, ,
\end{equation}
initially derived by Zaslavskii \cite{zas} in another context.
\section{Conclusion}
We have determined the explicit expressions of the electrostatic potential
and self-energy in the Schwarzschild black hole pierced by a cosmic string.
We can extend our method to the static, spherically symmetric spacetimes
pierced by a cosmic string when the electrostatic potential is known
in absence of a cosmic string: Brans-Dicke \cite{lin4}
and Reissner-Nordstr\"{o}m \cite{lea2}.
We have found again the upper entropy bound for a charged object
by employing thermodynamics of the Reissner-Nordstr\"{o}m black hole
pierced by a cosmic string. To prove this, we have used the value of the
electrostatic self-energy at the horizon of the Schwarzschild black hole
pierced by a cosmic string. This result confirms the physical importance of
the electrostatic self-force.
\newpage
|
\section{Introduction}
In a previous study \cite{krgl} we provided formal expressions for N-nucleon
effective generators of the Poincar\'e group derived from a field theory.
The ten hermitian generators derived from standard hermitian Lagrangians
describing interacting fields were blockdiagonalized by one and the same
unitary transformation, such that the space of a fixed number of nucleons
is separated from the rest of the space. The existence proof we carried
through makes use of
a formal power series expansion in the coupling constant g.
In this article we want to apply this procedure to the case of a
one-nucleon subspace. We shall study the effective generators in that
space in the first nontrivial order ${\cal O}(g^2)$. As
has to be expected this implies the question of mass renormalization in
the Hamiltonian and the boost operators, which are the four generators
carrying interactions in the instant form of relativistic dynamics.
The question will be, how the mass renormalization will come about and
whether after that step the Poincar\'e algebra for those effective
generators will still be valid.
We investigate these problems with the help of a Lagrangian describing
interacting scalar "nucleons" and mesons. In Section \ref{conmassren}
we formulate
the steps taken within the unitary transformation leading to the
effective generators. The mass renormalization in the effective
one-nucleon Hamiltonian is carried through in Section
\ref{massrenormalisation}.
An important
but lengthy algebra demonstrating the momentum independence of the mass
shift is deferred to the Appendix. Our result gained in the Hamiltonian
formalism is identical to the one gained by standard Feynman
methods in a manifestly covariant manner. This is shown in Section
\ref{compfeyn}.
The renormalisation of the effective boost operators is carried
through in Section \ref{renbo}. Finally we summarize in section \ref{summary}.
\section{Conditions for mass renormalisation}\label{conmassren}
Poincar\'e invariance requires that there exists a unitary representation of
the Poincar\'e group defined in a Hilbert space. The corresponding ten
generators fulfill the following set of commutation relations
\begin{eqnarray}
&&[{ P}_{i},H]=0\label{lie1}\\
&&[{ J}_{i},H]=0\label{lie2}\\
&&[{ P}_{i},{ P}_{j}]=0\label{lie3}\\
&&[{ J}_{i},{ J}_{j}]=i\epsilon_{ijk}{ J}_{k}\label{lie4}\\
&&[{ J}_{i},{ P}_{j}]=i\epsilon_{ijk}{ P}_{k}\label{lie5}\\
&&[{ J}_{i},{ K}_{j}]=i\epsilon_{ijk}{ K}_{k}\label{lie6}\\
&&[H,{ K}_{i}]=-i { P}_{i}\label{lie7}\\
&&[{ K}_{i},{ K}_{j}]=-i\epsilon_{ijk}{ J}_{k}\label{lie8}\\
&&[{ P}_{i},{ K}_{j}]=-i\delta_{ij}H\label{lie9}
\end{eqnarray}
Here $H$ is the Hamilton operator, $P_i$, and $J_i$ the three components
of the momentum and angular momentum operators and $K_i$ the three
boost operators. We note that in the instant form the four operators
$H$ and $K_i$ carry interactions. From Eqs. (\ref{lie1}) and (\ref{lie3})
we see that $H$ and $P_i$ can have simultaneous eigenstates which we
denote as $|\Psi_{\Fett p}\rangle$:
\begin{eqnarray}
H|\Psi_{\Fett p}\rangle&=&E|\Psi_{\Fett p}\rangle\label{heigen}\\
P_i|\Psi_{\Fett p}\rangle&=&p_i|\Psi_{\Fett p}\rangle\label{peigen}
\end{eqnarray}
As is well known it follows from Eqs. (\ref{lie7}) and (\ref{lie9}) that the
eigenvalues $E$ and $p_i$ are not independent, in fact one has
\begin{equation}
E=\sqrt{m^2+{\fett p}^2}\label{energie}
\end{equation}
Here $m$ is the physical rest mass defined in a system where $\fett p=0$.
Thus with $|\Psi_{\Fett 0}\rangle\equiv|\Psi_{\Fett p{\scriptscriptstyle=}\Fett 0}\rangle$
one has
\begin{eqnarray}
H|\Psi_{\Fett 0}\rangle&=&m|\Psi_{\Fett 0}\rangle\\
P_i|\Psi_{\Fett 0}\rangle&=&0
\end{eqnarray}
Now we would like to investigate a realisation in the following form.
We look at a system of two interacting scalar fields $\Psi$ and
$\Phi$ representing nucleons and mesons,
$\Psi$ being charged and $\Phi$ being a real field. The
interaction we look at is given by
\begin{equation}
{\cal L}_I(x)=g\Psi^\dagger(x)\Psi(x)\Phi(x)\label{lagrange}
\end{equation}
where
\begin{eqnarray}
\Psi(x)&=&{1\over\sqrt{2\pi}^3}\int \hskip-4pt d^3 q{1\over\sqrt{2E_{\Fett q}}}
\left(a_{\Fett q} {\rm e}^{-i qx}+b_{\Fett q}^\dagger{\rm e}^{i
qx}\right)\label{fieldpsi}\\
\Phi(x)&=&{1\over\sqrt{2\pi}^3}\int \hskip-4pt d^3 k{1\over\sqrt{2\omega_{\Fett k}}}
\left(c_{\Fett k} {\rm e}^{-i kx}+c_{\Fett k}^\dagger{\rm e}^{i
kx}\right)\label{fieldphi}\\
E_{\Fett q}&\equiv&\sqrt{m_0^2+\fett q^2}\label{ep}\\
\omega_{\Fett k}&\equiv&\sqrt{\mu_0^2+\fett k^2}\label{omegak}
\end{eqnarray}
Here $m_0$ and $\mu_0$ are the bare masses of nucleons and mesons,
$a^\dagger_{\Fett q}$ creates a nucleon while $b^\dagger_{\Fett q}$
creates an anti-nucleon and the $c_{\Fett k}$ refer to the mesons.
From Eq. (\ref{lagrange}) and with the help of the usual Noether
current arguments one can derive
expressions for the ten generators
of Poincar\'e transformations in terms of the field operators given in Eqs.
(\ref{fieldpsi}) and (\ref{fieldphi}) and performing space integrals
in terms of the particle creation and
annihilation operators.
In other words we want
to use the instant form proposed by Dirac \cite{dirac}.
Due to the usual equal time commutation relations of the fields
(\ref{fieldpsi}) and (\ref{fieldphi}) these generators fulfill the
Poincar\'e algebra (\ref{lie1})-(\ref{lie9}).
The states $|\Psi_{\Fett p}\rangle$ in Eqs. (\ref{heigen}) and
(\ref{peigen}) are now the physical eigenstates to the field theoretical
model Hamiltonian and total momentum operator. Clearly $H$ will depend
now on the bare masses $m_0$ and $\mu_0$. To calculate the physical mass
$m$ of one nucleon in terms of the bare masses $m_0$, $\mu_0$, and
the bare coupling constant $g$
we want to apply the Okubo transformation \cite{okubo}.
We regard the field theoretical ten generators to be presented as matrices in
the Fock space:
\begin{equation}
G=G\sum\limits_F|F\rangle\langle F|=\sum\limits_F|F\rangle\langle F|G
\label{matann}
\end{equation}
where $\{|F\rangle\}$ is a complete set of Fock space states and $G$
is any of the ten generators of the Poincar\'e group. Corresponding to
Eq. (\ref{matann}) we introduce Fock space matrix representations of the
generators which we call again $G$. As was shown in \cite{krgl} those
matrices fulfill the set (\ref{lie1}) - (\ref{lie9}). This matrix
representation of the algebra is our starting point.
As was proposed in \cite{glmu} and demonstrated in \cite{krgl}
the Okubo transformation carried through with one and the
same unitary matrix $\cal U$ block diagonalises the matrices of
all ten generators with
respect to two subspaces of the full Fock space. As a consequence the Lie
algebra of the Poincar\'e group given in Eqs. (\ref{lie1}) -
(\ref{lie9}) is still valid on both of the subspaces separately and we
want to call the projections of the transformed generators on either
of the subspaces "effective generators".
At this point we make the choice that one
of the two sub spaces consists of one nucleon only. Accordingly we define the
projection operators
\begin{equation}
\eta\equiv\int \hskip-4pt d^3 p\ a_{\Fett p}^\dagger|0\rangle\langle0|a_{\Fett p}
\label{eta}
\end{equation}
and
\begin{equation}
\Lambda\equiv 1-\eta
\end{equation}
In accordance to the definition of $\eta$ in Eq. (\ref{eta}) we
think of $\Lambda$ as being spanned by the set of eigenstates of the
three generators
$P_i$ which are not lying in the $\eta$ space.
Following Eq. (\ref{matann}) we write
\begin{equation}
G=G(\eta+\Lambda)=(\eta+\Lambda)G\label{matgen}
\end{equation}
Then under the action of the Okubo transformation we get
for the four generators $G\equiv H, K_i$
\begin{equation}
G{\buildrel{\cal U}\over\longrightarrow} {\cal U}G{\cal
U}^\dagger\equiv G'\label{genint}
\end{equation}
where $G'$ is now block diagonal:
\begin{equation}
G'=\eta G'\eta+\Lambda G'\Lambda
\end{equation}
Since we are working in the instant form, the generators $P_i$ and $J_i$ are
already diagonal in the Fock space and remain unchanged under $\cal U$
\cite{glmu}.
\begin{eqnarray}
G&{\buildrel{\cal U}\over\longrightarrow}& G\label{genfree}\\
G&=&P_i, J_i
\end{eqnarray}
Now we transform the matrix eigenvalue equations
\begin{eqnarray}
\langle F|H(\eta+\Lambda)|\Psi_{\Fett p}\rangle&=&E\langle
F|\Psi_{\Fett p}\rangle\\
\langle F|P_i(\eta+\Lambda)|\Psi_{\Fett p}\rangle&=&p_i\langle
F|\Psi_{\Fett p}\rangle
\end{eqnarray}
and get
\begin{eqnarray}
\langle F|H'(\eta+\Lambda)|\Psi'_{\Fett p}\rangle
&=&E\langle F|\Psi'_{\Fett p}\rangle\label{energieeff}\\
\langle F|P_i(\eta+\Lambda)|\Psi'_{\Fett p}\rangle
&=&p_i\langle F|\Psi'_{\Fett p}\rangle\label{impulseff}
\end{eqnarray}
with the transformed eigen state
\begin{equation}
\langle F|\Psi'_{\Fett p}\rangle\equiv\langle F|
{\cal U}(\eta+\Lambda)|\Psi_{\Fett p}\rangle
\end{equation}
Note that the solution vector $|\Psi_{\Fett p}\rangle$ is not
normalisable but the individual components $\langle F|\Psi_{\Fett p}\rangle$
exist, see for instance
\cite{textbook}.
The new set of generators $G'$ is unitarily equivalent to the original
one so that the eigen value problem presented in Eqs. (\ref{energieeff}) and
(\ref{impulseff}) describes the same physics as the set (\ref{heigen})
and (\ref{peigen}). However in contrast to the original problem (\ref{heigen})
and (\ref{peigen}) we can now focus on solutions $|\phi'_{\Fett p}\rangle$
to (\ref{energieeff}) and
(\ref{impulseff}) which are only lying in the $\eta$ space. These
solutions describe a single freely moving nucleon undergoing
self interactions due to the interaction part in the Lagrangian we
started with given in Eq. (\ref{lagrange}).
Since $\eta$ is the space of one nucleon only and because of
$P_i'=P_i$ we see that $|\phi'_{\Fett p}\rangle$
has to be a momentum eigenstate. One has
\begin{equation}
|\phi'_{\Fett p}\rangle=a^\dagger_{\Fett p}|0\rangle\equiv|{\fett p}\rangle
\end{equation}
Due to the coupling of nucleons and mesons occurring in
the original Hamiltonian the
transformed Hamilton operator $H'$ will carry an interaction:
\begin{equation}
\eta H'\eta=\eta H'\equiv\eta H_0 +\eta H'^{\phantom{i}\mbox{\tiny int}}
\label{hefform}
\end{equation}
where
\begin{equation}
\eta H_0\equiv\int \hskip-4pt d^3 p\ E_{\Fett p}a_{\Fett
p}^\dagger|0\rangle\langle0|a_{\Fett p}\label{h0}
\end{equation}
and $E_{\Fett p}$ being defined in Eq. (\ref{ep}) in terms of the bare
mass.
As we shall show in section \ref{massrenormalisation} $H'$ contains an
infinite constant which is related to the vacuum. Without that
constant the eigenvalue equation (\ref{energieeff})
(for an $\langle F|$ which is lying in the $\eta$ space) has to have the form
\begin{equation}
\left(\eta H_0 +\eta
H_{\mbox{\tiny nv}}^{'\mbox{\tiny int}}\right)|{\fett p}\rangle
{\buildrel !\over=}\sqrt{m^2+{\fett p}^2}|{\fett p}\rangle\label{start}
\end{equation}
which can be found by inserting Eqs. (\ref{energie}) and
(\ref{hefform})
into Eq. (\ref{energieeff}).
The effective potential carries now the index "nv" (no vacuum) as we
left out the vacuum contribution mentioned above.
Note that the energy eigenvalue $E$ given in Eq. (\ref{energie})
is unchanged under the unitary
transformation $\cal U$ and therefore
the physical
mass enters the square root
in Eq. (\ref{start}) unlike in the definition of $E_{\Fett q}$
in Eq. (\ref{ep}).
As a consequence of Eq. (\ref{start}) the effective Hamilton operator
without vacuum terms should have the form:
\begin{equation}
\eta H'_{\mbox{\tiny nv}}{\buildrel !\over=}\int \hskip-4pt d^3 p\ \sqrt{m^2+\fett p^2}
a_{\Fett p}^\dagger|0\rangle\langle0|a_{\Fett p}\label{expect}
\end{equation}
Since we are in principle able to calculate
$\eta H_{\mbox{\tiny nv}}^{'\mbox{\tiny int}}$
we can think of Eq. (\ref{start}) as defining $m$ as a function
of the three initial parameters
$g$, $m_0$, and $\mu_0$.
It is interesting to see how $\eta H_{\mbox{\tiny nv}}^{'\mbox{\tiny
int}}$ will replace $m_0$ in $H_0$ by the physical mass $m$. This will
be investigated in lowest non trivial order in the next section.
\section{Mass renormalisation}\label{massrenormalisation}
We start with the derivation of $H'$. According to \cite{okubo} and
\cite{krgl} the Okubo transformation (\ref{genint}) leads to
\begin{equation}
\eta H'=(1+A^\dagger A)^{^{-{1\over2}}}(\eta+A^\dagger)H
(\eta+A)(1+A^\dagger A)^{^{-{1\over2}}}\label{eff_form}
\end{equation}
where the operator $A$ has to be of the form
\begin{equation}
A\equiv\Lambda A\eta\label{aform}
\end{equation}
and where the block diagonalisation is guaranteed by the decoupling
equation
\begin{equation}
\Lambda \biggl([H,A]+H_I-AH_IA\biggr)\eta
= 0\label{bestH}
\end{equation}
Here we split the original $H$ into
\begin{equation}
H=H_0+H_I
\end{equation}
We solve the non linear equation (\ref{bestH}) perturbatively in
powers of the coupling constant $g$:
\begin{equation}
A=\sum\limits_{\nu=1}^\infty A_\nu\label{reihe1}
\end{equation}
where $A_\nu$ is proportional to $g^\nu$.
In order to calculate $H'$ up to second order in $g$ it is sufficient
to find $A_1$ as follows from Eq. (\ref{eff_form}):
\begin{eqnarray}
\eta H'&=&(1-\mbox{\large${1\over2}$} A^\dagger A)(\eta+A^\dagger)H
(\eta+A)(1-\mbox{\large${1\over2}$} A^\dagger A)+{\cal O}(g^3)\label{schrittheff}\\
&=&(1-\mbox{\large${1\over2}$} A_1^\dagger A_1)
(\eta+A_1^\dagger)H(\eta+A_1)(1-\mbox{\large${1\over2}$} A_1^\dagger A_1)
+{\cal O}(g^3)\label{heff}
\end{eqnarray}
In Eq. (\ref{heff}) we used the property (\ref{aform}) of $A$
and also $\eta H_0\Lambda=0=\Lambda H_0\eta$.
Due to Eqs. (\ref{bestH}) and (\ref{reihe1}) one way of finding $A_1$
is to solve the following commutator equation which follows from our
perturbation theoretical treatment of Eq. (\ref{bestH}):
\begin{equation}
{[}H_0,A_1]=-\Lambda H_I\eta\label{bestH1}
\end{equation}
So we need to know $H_I$. In the instant form the
interaction part of the Hamilton operator according to our model given
in Eq. (\ref{lagrange}) reads:
\begin{eqnarray}
H_I=-{g\over\sqrt {2\pi}^3}\int \hskip-4pt d^3 p\ \hskip-4pt d^3 q\ \hskip-4pt d^3 k\ &&{1\over\sqrt{8E_{\Fett
p}E_{\Fett q}\omega_{\Fett k}}}\nonumber\\
&&\times\biggl[a_{\Fett p}^\dagger a_{\Fett q}c_{\Fett k}^\dagger\delta^3(\fett
p-\fett q+\fett k)
+a_{\Fett p}^\dagger b_{\Fett q}^\dagger c_{\Fett k}^\dagger
\delta^3(\fett p+\fett q+\fett k)\nonumber\\
&&+a_{\Fett p}^\dagger a_{\Fett q}c_{\Fett k}\delta^3(\fett
p-\fett q-\fett k)
+a_{\Fett p}^\dagger b_{\Fett q}^\dagger c_{\Fett k}
\delta^3(\fett p+\fett q-\fett k)\nonumber\\
&&+b_{\Fett p}^\dagger b_{\Fett q}c_{\Fett k}^\dagger\delta^3(\fett
p-\fett q+\fett k)
+b_{\Fett p}a_{\Fett q} c_{\Fett k}^\dagger
\delta^3(-\fett p-\fett q+\fett k)\nonumber\\
&&+b_{\Fett p}^\dagger b_{\Fett q}c_{\Fett k}\delta^3(\fett
p-\fett q-\fett k)
+b_{\Fett p} a_{\Fett q} c_{\Fett k}
\delta^3(-\fett p-\fett q-\fett k)\biggr]\label{hint}
\end{eqnarray}
and one finds
\begin{eqnarray}
A_1&=&{g\over\sqrt{2\pi}^3}\int \hskip-4pt d^3 p\ \hskip-4pt d^3 q\ \hskip-4pt d^3 r\ \hskip-4pt d^3 k\
a^\dagger_{\Fett p}a^\dagger_{\Fett r}b^\dagger_{\Fett q}
c^\dagger_{\Fett k}|0\rangle\langle0|a_{\Fett r}
{\delta^3(\fett p+\fett q+\fett k)\over
\sqrt{8E_{\Fett p}E_{\Fett q}\omega_{\Fett k}}
(E_{\Fett p}+E_{\Fett q}+\omega_{\Fett k})}\nonumber\\
&&+{g\over\sqrt{2\pi}^3}\int \hskip-4pt d^3 p\ \hskip-4pt d^3 q\ \hskip-4pt d^3 k\ a_{\Fett p}^\dagger
c_{\Fett k}^\dagger|0\rangle\langle0|a_{\Fett q}{\delta^3(\fett
p-\fett q+\fett k)\over\sqrt{8E_{\Fett p}E_{\Fett q}\omega_{\Fett
k}}(E_{\Fett p}-E_{\Fett q}+\omega_{\Fett k})}\label{a1}
\end{eqnarray}
For the interaction (\ref{hint}) and for
$\eta$ projecting on the space of one nucleon
there is $\eta H_I\eta=0$ such that Eq. (\ref{heff})
simplifies to give:
\begin{eqnarray}
\eta H'&=&\eta H_0\eta+A_1^\dagger H_I+H_IA_1+A_1^\dagger H_0A_1
-\mbox{\large${1\over2}$} A_1^\dagger A_1H_0-\mbox{\large${1\over2}$} H_0A_1^\dagger A_1\\
&=&\eta H_0+\eta\left(\mbox{\large${1\over2}$} H_IA_1+\mbox{\large${1\over2}$} A_1^\dagger
H_I\right)\eta\label{heff1}
\end{eqnarray}
In the last step (\ref{heff1}) we used Eq. (\ref{bestH1}) and $\eta
H_0=\eta H_0\eta$ which follows from $H_0$ being a free operator. Using
Eqs. (\ref{hint}) and (\ref{a1}) we find
\begin{eqnarray}
\eta H_IA_1\eta=\! -{g^2\over8(2\pi)^3}&&\biggl\{\!\int \hskip-4pt d^3 p\ \hskip-4pt d^3 q\
a^\dagger_{\Fett p}|0\rangle\langle0|a_{\Fett p}
{1\over E_{\Fett p}E_{\Fett q}\omega_{\Fett p{\scriptscriptstyle +}\Fett q}}
\Bigl(
{1\over\omega_{\Fett p{\scriptscriptstyle +}\Fett q}+E_{\Fett p}+E_{\Fett q}}
+{1\over\omega_{\Fett p{\scriptscriptstyle +}\Fett q}-E_{\Fett
p}+E_{\Fett q}}
\Bigr)\nonumber\\
&&+\eta\int \hskip-4pt d^3 q\ \hskip-4pt d^3 p\ \hskip-4pt d^3 k\ {\delta^3(\fett p+\fett q+\fett k)
\delta^3(\fett p+\fett q+\fett k)\over E_{\Fett p}E_{\Fett
q}\omega_{\Fett k}}{1\over E_{\Fett p}+E_{\Fett q}+\omega_{\Fett
k}}\biggr\}
\nonumber\\
=-{g^2\over8(2\pi)^3}&&\int \hskip-4pt d^3 p\ \hskip-4pt d^3 q\
a^\dagger_{\Fett p}|0\rangle\langle0|a_{\Fett p}
{1\over E_{\Fett p}E_{\Fett q}\omega_{\Fett p{\scriptscriptstyle +}\Fett q}}
\Bigl(
{1\over\omega_{\Fett p{\scriptscriptstyle +}\Fett q}+E_{\Fett p}+E_{\Fett q}}
+{1\over\omega_{\Fett p{\scriptscriptstyle +}\Fett q}-E_{\Fett
p}+E_{\Fett q}}
\Bigr)\nonumber\\
&&+\eta\mbox{\it VAC}\label{ha}
\end{eqnarray}
The second term in Eq. (\ref{ha}) is an infinite constant
multiplied with the projector $\eta$. Since this constant
does not depend on quantum numbers of the nucleon we called it {\it
VAC} and omit it in what follows. Without this constant {\it
VAC} and with help of Eqs. (\ref{h0}) and (\ref{heff1}) we get
\begin{eqnarray}
\eta H'_{\mbox{\tiny nv}}&=&\int \hskip-4pt d^3 p\ a_{\Fett
p}^\dagger|0\rangle\langle0|a_{\Fett p}\nonumber\\
&&\times
\biggl\{E_{\Fett p}-{g^2\over8E_{\Fett p}(2\pi)^3}\int \hskip-4pt d^3 q\
{1\over E_{\Fett q}\omega_{\Fett p{\scriptscriptstyle +}\Fett q}}
\Bigl(
{1\over\omega_{\Fett p{\scriptscriptstyle +}\Fett q}+E_{\Fett p}+E_{\Fett q}}
+{1\over\omega_{\Fett p{\scriptscriptstyle +}\Fett q}-E_{\Fett
p}+E_{\Fett q}}
\Bigr)\biggr\}\nonumber\\
&&+{\cal O}(g^3)\label{hefferg}
\end{eqnarray}
Here we again put an index "nv" to indicate that vacuum terms have been
removed from this operator. We also note that the last denominator
cannot vanish for any choice of $\fett p$ and $\fett q$ since energy
and momentum conservation is impossible on a single
vertex described by our Lagrangian presented in Eq. (\ref{lagrange}).
Our task is to verify that this result is of the form of
Eq. (\ref{expect}) and to give a mathematical expression for the
physical mass $m$ in terms of $m_0$, $\mu_0$, and $g$.
To do so we introduce an expansion of $m$ into powers of the coupling
constant $g$:
\begin{equation}
m=m_0+\Delta m=m_0+\sum\limits_{\nu=1}^\infty m_\nu\label{massreihe}\\
\end{equation}
Similar to our notation in Eq. (\ref{reihe1}) the contributions $m_\nu$ are
proportional to $g^\nu$.
This ansatz ensures that in the limit $g\to0$ the physical and
the bare masses are equal. We note that the terms $m_\nu$ are
independent of momenta $\fett p$, of course.
A Taylor expansion of $\sqrt{m^2+{\fett p}^2}$ around $m=m_0$ gives:
\begin{equation}
\sqrt{m^2+\fett p^2}=E_{\Fett p}+{m_0\over E_{\Fett p}} (m_1+m_2)
+{\fett p^2\over2E_{\Fett p}^3}m_1^2+{\cal
O}(g^3)\label{taylor}
\end{equation}
We insert this expression into Eq. (\ref{expect}) and
find expressions for $m_1$ and $m_2$ by comparing equal powers of $g$
in Eqs. (\ref{expect}) and (\ref{hefferg}). We note that in
Eq. (\ref{hefferg}) there is
no contribution being linear in $g$, hence we get:
\begin{equation}
m_1=0
\end{equation}
We insert this into Eq. (\ref{taylor}) and find
\begin{equation}
\sqrt{m^2+{\fett p}^2}=E_{\Fett p}+{m_0\over E_{\Fett p}}m_2+{\cal
O}(g^3)\label{taylor1}
\end{equation}
Comparing Eq. (\ref{expect}) with Eq. (\ref{hefferg}) again and using
Eq. (\ref{taylor1}) this time we find:
\begin{equation}
m_2=-{g^2\over8m_0(2\pi)^3}\int \hskip-4pt d^3 q\
{1\over E_{\Fett q}\omega_{\Fett p{\scriptscriptstyle +}\Fett q}}
\Bigl(
{1\over\omega_{\Fett p{\scriptscriptstyle +}\Fett q}+E_{\Fett p}+E_{\Fett q}}
+{1\over\omega_{\Fett p{\scriptscriptstyle +}\Fett q}-E_{\Fett
p}+E_{\Fett q}}
\Bigr)\label{m2}
\end{equation}
Unfortunately, since the integrand
is clearly dependent on the parameter $\fett p$,
this result seems to violate our crucial assumption that
$m_2$ must not depend on external momenta $\fett p$. We also would like
to see consistency of our result with the well know expression for
the mass correction found by Feynman techniques which is of course
independent of initial momenta. For that purpose we look at the result
obtained from Feynman diagrams and we will show in the following
section that this result can indeed be proved to equal ours given by
Eq. (\ref{m2}) which verifies implicitly that Eq. (\ref{m2}) is not
dependent on $\fett p$. A direct and much harder
way to prove the latter
statement is presented in the Appendix.
Summarizing we have shown that indeed Eq. (\ref{expect}) holds in second
order in the coupling constant.
\section{comparison to mass renormalisation using Feynman
methods}\label{compfeyn}
We start with the same field theory as specified by
Eqs. (\ref{lagrange}), (\ref{fieldpsi}), and (\ref{fieldphi}). Also we
note that we often use covariant squares $q^2\equiv q^\mu q_\mu$ due to the
covariant Feynman formalism from now on in contrast to
the three dimensional vector products ${\fett q}^2$ of the last section. In
momentum space the Feynman propagators for the charged "nucleon" and the
uncharged meson are given by:
\begin{eqnarray}
i G(p)&\equiv&{i\over p^2-m_0^2+i\epsilon}\label{ferm}\\
i D(k)&\equiv&{i\over k^2-\mu_0^2+i\epsilon}\label{bos}
\end{eqnarray}
So we get the full propagator up to second order in the coupling
constant as
\begin{eqnarray}
i{\cal G}(p)&=&{i\over p^2-m_0^2+i\epsilon}+
{i\over p^2-m_0^2+i\epsilon}
i\Sigma(p){i\over p^2-m_0^2+i\epsilon}+{\cal O}(g^4)\label{fullprop1}
\\
i\Sigma(p)&\equiv&-{g^2\over(2\pi)^4}\int d^4\, q\ i G(q-p)i
D(q)\label{sigmadef}
\end{eqnarray}
Since $\Sigma(p)$ is a scalar it is actually a function $\Sigma(p^2)$. In a
well known manner \cite{mandl}
using the requirement that ${\cal G}(p)$ has a pole at the
physical mass $m$ one finds the mass shift given by
\begin{eqnarray}
m_2^{\mbox{\tiny F}}&=&-{1\over2m_0}\Sigma(m^2)\nonumber\\
&=&-{1\over2m_0}\Sigma(m_0^2)+{\cal O}(g^4)
\label{m2feyn}
\end{eqnarray}
The suffix 'F' denotes that we got this result by using Feynman
diagrams.
By construction it is guaranteed this time that $m_2^{\mbox{\tiny F}}$
is not dependent on initial momenta.
We carry out the $q_0$ integration in Eq. (\ref{sigmadef}) in
order to get a three dimensional form:
\begin{eqnarray}
m_2^{\mbox{\tiny F}}&=&{i g^2\over2(2\pi)^4m_0}\int \hskip-4pt d^3 q\
\int\limits_{-\infty}^\infty d\, q_0\ {1\over\left(q_0-E_{\Fett p}\right)^2
-\left(\fett q-\fett p\right)^2-m_0^2+i\epsilon}{1\over q_0^2-\fett
q^2-\mu_0^2+i\epsilon}\nonumber\\
&=&{i g^2\over2(2\pi)^4m_0}\int \hskip-4pt d^3 q\
\int\limits_{-\infty}^\infty d\, q_0\ {1\over\left(q_0-E_{\Fett p}\right)^2
-E^2_{\Fett q{\scriptscriptstyle -}\Fett p}+i\epsilon}
{1\over q_0^2-\omega^2_{\Fett q}+i\epsilon}
\label{step3}
\end{eqnarray}
Here the terms $E_{\Fett p}$ and $E_{\Fett q{\scriptscriptstyle -}\Fett p}$
occur as we calculate $\Sigma(p^2=m_0^2)$.
A simple integration yields
\begin{eqnarray}
m_2^{\mbox{\tiny F}}&=&
{i^2g^2\over2m_0(2\pi)^3}\int \hskip-4pt d^3 q \Biggl[
{1\over-2E_{\Fett q{\scriptscriptstyle-}\Fett p}}
\,\,{1\over\left(E_{\Fett p}-E_{\Fett q{\scriptscriptstyle-}\Fett p}
\right)^2-\omega_{\Fett q}^2}\nonumber\\
&&\qquad\qquad\qquad\qquad\qquad\qquad
+{1\over\left(E_{\Fett p}+\omega_{\Fett q}\right)^2
-E^2_{\Fett q{\scriptscriptstyle-}\Fett p}}\,\,
{1\over-2\omega_{\Fett q}}\Biggr]\label{step4}
\end{eqnarray}
For the same reason as stated above the denominators cannot vanish.
We go ahead adding zeros in the numerators:
\begin{eqnarray}
m_2^{\mbox{\tiny F}}&=&
{g^2\over4m_0(2\pi)^3}\int \hskip-4pt d^3 q\ {1\over E_{\Fett
q\scriptscriptstyle-\Fett p}\omega_{\Fett q}}
\Biggl[
\mbox{\large${1\over2}$}{2\omega_{\Fett q}+E_{\Fett q\scriptscriptstyle-\Fett p}-E_{\Fett p}
-E_{\Fett q\scriptscriptstyle-\Fett p}+E_{\Fett p}\over
\left(E_{\Fett p}-E_{\Fett q\scriptscriptstyle-\Fett p}-\omega_{\Fett q}
\right)
\left(E_{\Fett p}-E_{\Fett q\scriptscriptstyle-\Fett p}+\omega_{\Fett q}
\right)}\nonumber\\
&&+\mbox{\large${1\over2}$}{2E_{\Fett q\scriptscriptstyle-\Fett p}-E_{\Fett p}-\omega_{\Fett q}
+E_{\Fett p}+\omega_{\Fett q}\over
\left(E_{\Fett p}+\omega_{\Fett q}-E_{\Fett q\scriptscriptstyle-
\Fett p}\right)
\left(E_{\Fett p}+\omega_{\Fett q}+E_{\Fett q\scriptscriptstyle-
\Fett p}\right)}\Biggr]\label{step6}
\end{eqnarray}
and find
\begin{equation}
m_2^{\mbox{\tiny F}}=
{g^2\over8m_0(2\pi)^3}
\int \hskip-4pt d^3 q\ {1\over E_{\Fett q\scriptscriptstyle-\Fett p}\omega_{\Fett
q}}
\Biggl[
{1\over E_{\Fett p}-E_{\Fett q\scriptscriptstyle-\Fett p}
-\omega_{\Fett q}}
-{1\over E_{\Fett p}+E_{\Fett q\scriptscriptstyle-\Fett p}
+\omega_{\Fett q}}
\Biggr]\label{step7}
\end{equation}
Now we see after the substitution $\fett q\to\fett q+\fett p$
that $m_2^{\mbox{\tiny F}}$ is equal to $m_2$ given in Eq. (\ref{m2}).
Apparently in that form (\ref{step7}) or (\ref{m2}) it is far from obvious
that the integral does not depend on $\fett p$. Nevertheless it can be
shown to be true as presented in the Appendix. Also the steps
backwards from Eq. (\ref{step7}) to Eq. (\ref{step3}) are not obvious
without knowing the final result. However the three dimensional
formalism we are aiming at is much closer to the standard form of a
non relativistic Schr\"odinger equation used in nuclear physics than the
four dimensional off-the-mass-shell structures. Therefore we think
it is justified to put now efforts into three dimensional Hamiltonian
forms derived from field theory.
\section{renormalisation of the boost operators}\label{renbo}
In section \ref{massrenormalisation} we found that the transformed
Hamilton operator in the $\eta$ space, $\eta H'_{\mbox{\tiny{nv}}}$,
up to additive constants looks like the free Hamilton operator in the,
$\eta$ space $\eta H_0$, after replacing the bare nucleon mass $m_0$ by
the physical nucleon mass $m$. A similar situation
should hold true for the three
transformed boost operators, i.e. we expect that $\eta\subsub
Ki{\mbox{\tiny{nv}}}'$, $i=1,2,3$, are equal to $\eta\subsub K0i$
expressed in terms of $m$ instead of $m_0$.
One reason for expecting that is the
following: The solutions $|\phi'_{\Fett p}\rangle$ to the
equations (\ref{energieeff}) and (\ref{impulseff}) which lie in the
$\eta$ space represent a free
single nucleon with the physical mass $m$.
This is because we chose $\eta$ to be the space of a single nucleon.
As a consequence the boost operators $\eta\subsub
Ki{\mbox{\tiny{nv}}}'$corresponding to $\eta
H_{\mbox{\tiny{nv}}}'$ should also be free boost operators generating
boosts of a free particle with mass $m$.
Another reason can be addressed by looking at the Lie algebra of the
Poincar\'e group given in Eqs. (\ref{lie1})-(\ref{lie9}):
In section \ref{conmassren}
we pointed out that this Lie algebra is still valid on the $\eta$
space on its own. While we noted in section \ref{conmassren} that the
generators $P_i$ and $J_i$ remain unchanged under the action of an
Okubo transformation we proved in section \ref{massrenormalisation}
that the transformed Hamilton operator, up to constants, looks like a
free Hamilton operator where the bare mass $m_0$ is replaced by the
physical mass $m$. As the Lie algebra is fulfilled in the case of free
particles we conclude that the corresponding transformed
boost operators should also look like free boost operators expressed
in terms of $m$ instead of $m_0$.
More precisely:
While the three free boost operators in the $\eta$ space are given by
\begin{equation}
\eta\subsub K0i={i\over2}\eta\int \hskip-4pt d^3 p\ a_{\Fett p}^\dagger\Bigl(
\sqrt{m_0^2+{\fett p}^2}{\partial\over\partial p_i}
+{\partial\over\partial p_i}\sqrt{m_0^2+{\fett p}^2}\Bigr)
a_{\Fett p}\qquad i=1,2,3
\end{equation}
we expect the transformed full boost operators in the $\eta$ space,
after dropping constants related to the vacuum, to look like
\begin{equation}
\eta\subsub Ki{\mbox{\tiny nv}}'
{\buildrel!\over=}
{i\over2}\eta\int \hskip-4pt d^3 p\ a_{\Fett p}^\dagger\Bigl(\sqrt{m^2+{\fett p}^2}
{\partial\over\partial p_i}+{\partial\over\partial
p_i}\sqrt{m^2+{\fett p}^2}
\Bigr)
a_{\Fett p}\qquad i=1,2,3\label{kexp}
\end{equation}
Like in Eq. (\ref{hefform})
we put an index 'nv' which indicates that we did not take any
constants into account.
We are now going to verify that Eq. (\ref{kexp}) holds true indeed.
According to \cite{okubo} and \cite{krgl} one finds the following expression
for the transformed boost operators in the $\eta$ space after an Okubo
transformation:
\begin{equation}
\eta K_i'
=\eta \subsub K0i+\eta\left(\mbox{\large${1\over2}$} \subsub KIiA_1+\mbox{\large${1\over2}$} A_1^\dagger
\subsub KIi\right)\eta\label{keff1}
\end{equation}
where $A_1$ is given by Eq. (\ref{a1}).
This result corresponds to Eq. (\ref{heff1}).
To go ahead one
derives an expression for $\subsub KIi$ from
the interaction part of the Lagrangian given in Eq. (\ref{lagrange}):
\begin{eqnarray}
\subsub KIi=-i{g\over\sqrt {2\pi}^3}\int \hskip-4pt d^3 p\
\hskip-4pt d^3 q\ \hskip-4pt d^3 k\! &&{1\over\sqrt{8E_{\Fett
p}E_{\Fett q}\omega_{\Fett k}}}\nonumber\\
&&\times\biggl[a_{\Fett p}^\dagger a_{\Fett q}c_{\Fett k}^\dagger
{\partial\over\partial k_i}\delta^3(\fett
p-\fett q+\fett k)
+a_{\Fett p}^\dagger b_{\Fett q}^\dagger c_{\Fett k}^\dagger
{\partial\over\partial k_i}\delta^3(\fett p+\fett q+\fett k)\nonumber\\
&&-a_{\Fett p}^\dagger a_{\Fett q}c_{\Fett k}
{\partial\over\partial k_i}\delta^3(\fett
p-\fett q-\fett k)
-a_{\Fett p}^\dagger b_{\Fett q}^\dagger c_{\Fett k}
{\partial\over\partial k_i}\delta^3(\fett p+\fett q-\fett k)\nonumber\\
&&+b_{\Fett p}^\dagger b_{\Fett q}c_{\Fett k}^\dagger
{\partial\over\partial k_i}\delta^3(\fett
p-\fett q+\fett k)
+b_{\Fett p}a_{\Fett q} c_{\Fett k}^\dagger
{\partial\over\partial k_i}\delta^3(-\fett p-\fett q+\fett k)\nonumber\\
&&-b_{\Fett p}^\dagger b_{\Fett q}c_{\Fett k}
{\partial\over\partial k_i}\delta^3(\fett
p-\fett q-\fett k)
-b_{\Fett p} b_{\Fett q} c_{\Fett k}
{\partial\over\partial k_i}
\delta^3(-\fett p-\fett q-\fett k)\biggr]\label{kint}
\end{eqnarray}
Note the formal similarity to $H_I$ given in Eq. (\ref{hint}).
The expression (\ref{kint}) can be inserted into Eq. (\ref{keff1}).
Again it turns out that $K_i'$ contains an additive infinite
constant being related to the vacuum. After dropping this constant one
arrives at an intermediate result:
\begin{eqnarray}
\subsub Ki{\mbox{\tiny nv}}'
&=&\eta\subsub K0i\nonumber\\
&&-{i g^2\over16E_{\Fett p}(2\pi)^3}
\int \hskip-4pt d^3 p\ \hskip-4pt d^3 q\
{1\over E_{\Fett q}\omega_{\Fett p{\scriptscriptstyle +}\Fett q}}
\Bigl(
{1\over\omega_{\Fett p{\scriptscriptstyle +}\Fett q}+E_{\Fett p}+E_{\Fett q}}
+{1\over\omega_{\Fett p{\scriptscriptstyle +}\Fett q}+E_{\Fett
p}-E_{\Fett q}}
\Bigr)\nonumber\\
&&\times \biggl\{a_{\Fett
p}^\dagger|0\rangle\langle0|{\partial\over\partial p_i} a_{\Fett p}
-\biggl({\partial\over\partial p_i}a_{\Fett
p}^\dagger\biggr)|0\rangle\langle0|a_{\Fett p}\biggr\}
+{\cal O}(g^3)\label{keff2}
\end{eqnarray}
After rewriting $\eta\subsub K0i$ into
\begin{equation}
\eta\subsub K0i=\mbox{\large${i\over2}$}\int \hskip-4pt d^3 p\ E_{\Fett p}
\biggl(a_{\Fett p}^\dagger|0\rangle{\partial\over\partial
p_i}\langle0|a_{\Fett p}-\Bigl({\partial\over\partial p_i}
a_{\Fett p}^\dagger\Bigr)|0\rangle\langle0|a_{\Fett p}\biggr)\label{k0neu}
\end{equation}
we can modify Eq. (\ref{keff2}) to give
\begin{eqnarray}
\eta \subsub Ki{\mbox{\tiny nv}}'&=&\!\mbox{\large${i\over2}$}\!
\int \hskip-4pt d^3 p\
\biggl\{E_{\Fett p}-{g^2\over8E_{\Fett p}(2\pi)^3}\int \hskip-4pt d^3 q\
{1\over E_{\Fett q}\omega_{\Fett p{\scriptscriptstyle +}\Fett q}}
\Bigl(
{1\over\omega_{\Fett p{\scriptscriptstyle +}\Fett q}+E_{\Fett p}+E_{\Fett q}}
+{1\over\omega_{\Fett p{\scriptscriptstyle +}\Fett q}+E_{\Fett
p}-E_{\Fett q}}
\Bigr)\biggr\}\nonumber\\
&&\times\biggl(a_{\Fett p}^\dagger|0\rangle{\partial\over\partial
p_i}\langle0|a_{\Fett p}-\Bigl({\partial\over\partial p_i}
a_{\Fett p}^\dagger\Bigr)|0\rangle\langle0|a_{\Fett p}\biggr)
+{\cal O}(g^3)\label{kefferg}
\end{eqnarray}
We note that the curly bracket is identical to the one in Eq. (\ref{hefferg})
for $\eta H'_{\mbox{\tiny nv}}$.
As a consequence we see that the integral in $\fett q$ is closely related
to the definition of $m_2$ given in Eq. (\ref{m2}) and we can rewrite
Eq. (\ref{kefferg}) in terms of $m_2$.
\begin{equation}
\eta \subsub Ki{\mbox{\tiny nv}}'=\mbox{\large${i\over2}$}
\int \hskip-4pt d^3 p\
\biggl(E_{\Fett p}+{m_0\over E_{\Fett p}}m_2\biggr)
\biggl(a_{\Fett p}^\dagger|0\rangle{\partial\over\partial
p_i}\langle0|a_{\Fett p}-\Bigl({\partial\over\partial p_i}
a_{\Fett p}^\dagger\Bigr)|0\rangle\langle0|a_{\Fett p}\biggr)
+{\cal O}(g^3)\label{kefferg1}
\end{equation}
Now we make use of an intermediate result presented in
Eq. (\ref{taylor1}) and find as the desired result
\begin{equation}
\eta \subsub Ki{\mbox{\tiny nv}}'=\mbox{\large${i\over2}$}
\int \hskip-4pt d^3 p\
\biggl(\sqrt{m^2+{\fett p}^2}\biggr)
\biggl(a_{\Fett p}^\dagger|0\rangle{\partial\over\partial
p_i}\langle0|a_{\Fett p}-\Bigl({\partial\over\partial p_i}
a_{\Fett p}^\dagger\Bigr)|0\rangle\langle0|a_{\Fett p}\biggr)
+{\cal O}(g^3)\label{kefferg2}
\end{equation}
which proves that the three transformed boost operators $\eta \subsub
Ki{\mbox{\tiny nv}}'$ look like free boost operators where
the bare nucleon mass $m_0$ has been replaced by the physical mass
$m$. Then the Poincar\'e algebra is of course fulfilled.
\section{Summary}\label{summary}
Starting from a Lagrangian describing charged scalar "nucleons"
and uncharged mesons which interact via a simple vertex
expression the ten effective generators for Poincar\'e transformations for
one nucleon are derived in lowest nontrivial order in the coupling
constant. We used the Okubo transformation
The dynamics of this system of one nucleon is governed
by the generator for time translations, the effective
Hamilton operator. Looking at the simultaneous eigenstates to the
Hamilton and the three momentum operators general arguments lead to
the statement that the eigenvalues with respect to the Hamilton operator
have to be of the form $E=\sqrt{m^2+\fett{p}^2}$ where $p_i$ are the
eigenvalues to the three components of the momentum operator and $m$
is the total rest mass of the system being described. In our case that
system is the system of one single nucleon undergoing self interactions
due to the interaction part in the Lagrangian proposed above and so
the total rest mass of that system is the total rest mass of a
physical nucleon.
We used that last statement to derive an expression for that physical
mass in terms of the three initial parameters, the bare masses of the
nucleon and the meson $m_0$ and $\mu_0$ and the coupling constant
$g$. We did this in a perturbation theoretical manner and gave an
expression for $m$ in section \ref{massrenormalisation}
which is correct up to second order in the coupling
constant. After some
calculation we could show that
$m$, at least up to second order,
is not dependent on the momentum of the particle. This is in
contrast to a first look at the analytical expression for $m$.
In section
\ref{compfeyn} we compared this result to the expression found for $m$
using Feynman techniques and revealed that the two calculations lead to
the same result. The result gained by Feynman techniques
is strictly not dependent on the momentum
of the particle.
By explicit calculation in section \ref{renbo} we showed that
the three effective boost operators can be rewritten in terms of the
physical mass of the nucleon $m$, again we did this in second order in
the coupling constant only. The results are three free effective
boost operators
where the bare mass of the nucleon is just replaced by the physical
mass.
In the Appendix that question of the '$\fett p$
independence' of the mass posed in section \ref{compfeyn} was answered
without knowledge of the Feynman result by a lengthy but straight
forward calculation.
This completes our discussion of this model of one physical
nucleon stating that up to second order the effective
generators of Poincar\'e
transformations are equal to the well known free generators of the
Poincar\'e group except that in the Hamilton and the three boost
operators the bare mass of the nucleon is replaced by its physical
mass.
Clearly the next step will be to investigate two interacting nucleons.
In \cite{glmu} it has been shown that in leading order ${\cal O}(g^2)$
the effective generators fulfill the Poincar\'e algebra. But this does
not yet involve loop integrals and thus renormalizations. They will occur
in the order ${\cal O}(g^4)$. The interesting question will be, whether the
effective generators after renormalization (in the Hamiltonian formalism)
will retain their property formally found in \cite{krgl} to fulfill the
Poincar\'e algebra. This work is under investigation.
\section{Appendix}
In section \ref{compfeyn} we have seen that the second order
mass shift calculated by the
method of unitary transformation in the Hamilton formalism is equal to
the mass shift resulting from a calculation by Feynman methods. Hence
we have shown at the same time that our expression given in
Eq. (\ref{m2}) is not dependent on the initial momentum $\fett p$ which
is in contrast to a first glance at Eq. (\ref{m2}) where the integrand
is a function of $\fett p$.
Now a lengthy calculation which does not make use of results which
have been derived in the framework of a covariant formalism
shows that after the integration this dependence on $\fett p$ is lost.
This will be shown now.
First we notice that actually the integral occurring in Eq. (\ref{m2})
is ultra-violet divergent. We abbreviate and introduce a cutoff $\lambda$:
\begin{equation}
\int\limits_{|\Fett q|{\scriptscriptstyle\le\lambda}}
\hskip-4pt d^3 q\ F(\fett p,\fett q)\equiv
\int\limits_{|\Fett q|{\scriptscriptstyle\le\lambda}} \hskip-4pt d^3 q\
{1\over E_{\Fett q}\omega_{\Fett p{\scriptscriptstyle +}\Fett q}}
\Bigl(
{1\over\omega_{\Fett p{\scriptscriptstyle +}\Fett q}+E_{\Fett p}+E_{\Fett q}}
+{1\over\omega_{\Fett p{\scriptscriptstyle +}\Fett q}-E_{\Fett
p}+E_{\Fett q}}
\Bigr)\label{Cutoff}
\end{equation}
so that in the limit $\lambda\to\infty$ we regain the integral
presented in Eq. (\ref{m2}).
Now we split the convergent (regularized)
integral given in Eq. (\ref{Cutoff}):
\begin{equation}
\int\limits_{|\Fett q|{\scriptscriptstyle\le\lambda}} \hskip-4pt d^3 q\ F(\fett p,\fett q)
=\int\limits_{|\Fett q|{\scriptscriptstyle\le\lambda}} \hskip-4pt d^3 q\
\bigl(F(\fett p,\fett q)-F(0,\fett q)\bigr)
+\int\limits_{|\Fett q|{\scriptscriptstyle\le\lambda}} \hskip-4pt d^3 q\
F(0,\fett q)\label{split}
\end{equation}
It can be shown that the first integral is no longer ultra-violet
divergent as $\lambda\to\infty$:
\begin{equation}
\lim_{\lambda\to\infty}
\int\limits_{|\Fett q|{\scriptscriptstyle\le\lambda}} \hskip-4pt d^3 q\
\bigl(F(\fett p,\fett q)-F(0,\fett q)\bigr)<\infty\label{absch}
\end{equation}
We see that
the "$\fett p$ dependence" of $m_2$ is now
accessible to investigation since its origin is the first, well
defined expression in Eq. (\ref{split}). Consequently, to investigate
the $\fett p$ dependence of $\int \hskip-4pt d^3 q\ F(\fett q,\fett p)$ we now draw our
attention to the first term.
We can simplify the integrals analytically:
\begin{eqnarray}
\int\limits_{|\Fett q|{\scriptscriptstyle\le\lambda}}\!\!
\hskip-4pt d^3 q\ F(0,\fett q)&=&8\pi\int\limits_0^\lambda\!\! d\, q\
{q^2\over\sqrt{\mu_0^2+q^2}\sqrt{m_0^2+q^2}}
{\sqrt{\mu_0^2+q^2}+\sqrt{m_0^2+q^2}\over
(\sqrt{\mu_0^2+q^2}+\sqrt{m_0^2+q^2})^2-m_0^2}\label{f0}\\
\int\limits_{|\Fett q|{\scriptscriptstyle\le\lambda}}\!\!
\hskip-4pt d^3 q\ F(\fett p,\fett q)&=&
2\pi\!\!\int\limits_0^\lambda\!\! d\, q\
{q\over\sqrt{m_0^2+q^2}p}\ln
{\left(\sqrt{m_0^2+p^2}+\sqrt{\mu_0^2+(p+q)^2}\right)^2-m_0^2-p^2\over
\left(\sqrt{m_0^2+p^2}+\sqrt{\mu_0^2+(p-q)^2}\right)^2-m_0^2-p^2}
\label{fp}
\end{eqnarray}
where
\begin{eqnarray}
p&\equiv&|\fett p|\\
q&\equiv&|\fett q|
\end{eqnarray}
We introduce dimensionless quantities
\begin{eqnarray}
x&\equiv&{q\over m_0}\label{def00}\\
\alpha&\equiv&{p\over m_0}\label{def01}\\
\beta_0&\equiv&{\mu_0\over m_0}\label{def02}\\
\Lambda&\equiv&{\lambda\over p}\label{def03}
\end{eqnarray}
and define
\begin{eqnarray}
{\cal I}_\alpha^\Lambda(\beta)
\equiv2\pi\int\limits_0^\Lambda d\, x&&\Biggl\{
4{x^2\over\sqrt{1+x^2}\sqrt{\beta^2+x^2}}
{\sqrt{1+x^2}+\sqrt{\beta^2+x^2}\over
\left(\sqrt{1+x^2}+\sqrt{\beta^2+x^2}\right)^2-1}\nonumber\\
&&-{1\over\alpha}{x\over\sqrt{1+x^2}}\ln
{\left(\sqrt{1+x^2}+\sqrt{\beta^2+(x+\alpha)^2}\right)^2-1-\alpha^2\over
\left(\sqrt{1+x^2}+\sqrt{\beta^2+(x-\alpha)^2}-1-\alpha^2\right)^2}\Biggr\}
\label{term1}
\end{eqnarray}
Then find
\begin{equation}
\lim_{\Lambda\to\infty}{\cal I}_\alpha^\Lambda(\beta_0)=
\int \hskip-4pt d^3 q\ \left(F(0,\fett q)-F(\fett p,\fett q)\right)\label{ibeta0}
\end{equation}
For any value of $\alpha$ and $\Lambda$
we want to think of ${\cal I}_\alpha^\Lambda(\beta)$ as being an
analytical function dependent on the complex
parameter $\beta$. Due to Eq. (\ref{absch}) this also includes the
case $\Lambda=\infty$. From Eq. (\ref{ibeta0}) we see
that ${\cal I}_\alpha^\Lambda(\beta)$ becomes a physical
meaning at the point $\beta_0$ defined by
Eq. (\ref{def02}) and in the limit $\Lambda=\infty$.
For the case of $\lambda<\infty$
we want to modify the second term of Eq. (\ref{ibeta0}):
\begin{eqnarray}
\lefteqn{-\int\limits_{|\Fett q|{\scriptscriptstyle\le}\lambda} \hskip-4pt d^3 q\ F(\fett p,\fett q)}
\nonumber\\
&=&{-2\pi\over\alpha}\int\limits_0^\Lambda d\, x
{x\over\sqrt{1+x^2}}\ln
{\left(\sqrt{1+x^2}+\sqrt{\beta^2+(x+\alpha)^2}\right)^2-1-\alpha^2\over
\left(\sqrt{1+x^2}+\sqrt{\beta^2+(x-\alpha)^2}\right)^2-1-\alpha^2}
\nonumber\\
&=&{-2\pi\over\alpha}\int
\limits_{-\Lambda}^\Lambda d\, x
{x\over\sqrt{1+x^2}}\ln
\left(\left(\sqrt{1+x^2}+\sqrt{\beta^2+(x+\alpha)^2}\right)^2-1-\alpha^2
\right)\nonumber\\
&=&{-2\pi\over\alpha}\sqrt{1+x^2}\ln
\left\{\left(\sqrt{1+x^2}+\sqrt{\beta^2+(x+\alpha)^2}\right)^2-1-\alpha^2
\right\}\Biggl|_{-\Lambda}^\Lambda\\
&+&{2\pi\over\alpha}\int
\limits_{-\Lambda}^\Lambda d\, x\sqrt{1+x^2}
{2\left(\sqrt{1+x^2}+\sqrt{\beta^2+(x+\alpha)^2}\right)
\left({x\over\sqrt{1+x^2}}+{x+\alpha\over\sqrt{\beta^2+(x+\alpha)^2}}\right)
\over\left(\sqrt{1+x^2}+\sqrt{\beta^2+(x+\alpha)^2}\right)^2-1-\alpha^2}
\nonumber
\end{eqnarray}
Here the first of the two terms is equal to $-4\pi$ as $\Lambda$ goes to
infinity. So we reformulate Eq. (\ref{term1}) as:
\begin{eqnarray}
\lim_{\Lambda\to\infty}{\cal I}_\alpha^\Lambda(\beta)\nonumber\\
=-4\pi&&+2\pi\lim_{\Lambda\to\infty}
\int\limits_{-\Lambda}^\Lambda d\, x\Biggl\{
{2x^2\over\sqrt{1+x^2}\sqrt{\beta^2+x^2}}
{\sqrt{1+x^2}+\sqrt{\beta^2+x^2}\over
\left(\sqrt{1+x^2}+\sqrt{\beta^2+x^2}\right)^2-1}\nonumber\\
&&+{2\sqrt{1+x^2}\over\alpha}
{\left(\sqrt{1+x^2}+\sqrt{\beta^2+(x+\alpha)^2}\right)
\left({x\over\sqrt{1+x^2}}+{x+\alpha\over\sqrt{\beta^2+(x+\alpha)^2}}\right)
\over\left(\sqrt{1+x^2}+\sqrt{\beta^2+(x+\alpha)^2}\right)^2-1-\alpha^2}
\Biggr\}
\end{eqnarray}
It can be shown that no singularities are lying on the path of integration. To
solve this integral we introduce the following substitution:
\begin{equation}
\sqrt{x^2+1}\equiv z(x-i)
\label{subst}
\end{equation}
This implies that the path of integration is shifted into the upper
half of the complex
$z$-plane going from $-1$ to $+1$ on a half
circle. After the substitution the integrand is a rational function
which can be integrated analytically. We also want to note
that it is important
to treat the limits correctly. The
result is:
\begin{eqnarray}
&&\lim_{\Lambda\to\infty}{\cal I}_\alpha^\Lambda(\beta)\nonumber\\
&&=-4\pi+2\pi\lim_{\Lambda\to\infty}\Biggl[
-\beta^4\int\limits_
{\sqrt{{-\Lambda+\beta i\over-\Lambda-\beta i}}}^
{\sqrt{{\Lambda+\beta i\over\Lambda-\beta i}}}
d\, z{1\over z^2-1}
{i^2{(z^2+1)^2\over(z^2-1)^2}
\over
\left(i{z^2+1\over z^2-1}\beta-\zeta\right)
\left(i{z^2+1\over z^2-1}\beta+\zeta\right)
}\nonumber\\
&&+\int\limits_
{\sqrt{-\Lambda+i\over-\Lambda-i}}^
{\sqrt{\Lambda+i\over\Lambda-i}}
d\, z{1\over z^2-1}
{i^2{(z^2+1)^2\over(z^2-1)^2}(\beta^2-2)
\over
\left(i{z^2+1\over z^2-1}-\zeta\right)
\left(i{z^2+1\over z^2-1}+\zeta\right)
}
\nonumber\\
-&&{4\over\alpha}\int\limits_
{\sqrt{-\Lambda+i\over-\Lambda-i}}^
{\sqrt{\Lambda+i\over\Lambda-i}} d\, z
{z^2\over\left(z^2-1\right)^3}{-2i{z^2+1\over z^2-1}+\alpha(\beta^2-2)\over
\left(i{z^2+1\over z^2-1}-\eta+\alpha-\xi\right)
\left(i{z^2+1\over z^2-1}-\eta+\alpha+\xi\right)}\nonumber\\
-&&{4\over\alpha}\int\limits_
{\sqrt{-\Lambda+\alpha+\beta i\over-\Lambda+\alpha-\beta i}}^
{\sqrt{\Lambda+\alpha+\beta i\over\Lambda+\alpha-\beta i}} d\, z
{z^2\over\left(z^2-1\right)^3}
{1\over i\left({z^2+1\over z^2-1}-1\right)}
{1\over i\left({z^2+1\over z^2-1}+1\right)}
{1\over\left(i\beta{z^2+1\over z^2-1}-\eta-\xi\right)}
{1\over\left(i\beta{z^2+1\over z^2-1}-\eta+\xi\right)}
\nonumber\\
&&\phantom{AAAAAAAA}
\times\Biggl(2\alpha\beta^2+2\alpha^3\beta^2-\alpha\beta^4
+(-3\beta^3-5\alpha^2\beta^3+\beta^5)i{z^2+1\over z^2-1}
\nonumber\\
&&\phantom{AAAAAAAAAAAAA}
+3\alpha\beta^4i^2{(z^2+1)^2\over (z^2-1)^2}
-2\beta^3i^3{(z^2+1)^3\over (z^2-1)^3}\Biggr)
\Biggr]\label{integral}
\end{eqnarray}
where we used:
\begin{eqnarray}
\xi&\equiv&{\beta\over2}\sqrt{(\alpha^2+1)(\beta^2-4)}\label{def1}\\
\eta&\equiv&{\alpha\beta^2\over2}\label{def2}\\
\zeta&\equiv&{\beta\over2}\sqrt{\beta^2-4}\label{def3}
\end{eqnarray}
For $\beta\ge2$ we see that $\xi$ and $\zeta$ remain real and we want
to concentrate on that case first.
The integral Eq. (\ref{integral}) can be carried out analytically and
we find for $\beta\ge2$:
\begin{eqnarray}
\lefteqn{\lim_{\Lambda\to\infty}{\cal I}_\alpha^\Lambda(\beta)}\nonumber\\
&=&-2\pi\Biggl[
-{1\over\alpha}\sqrt{1+(\eta-\alpha+\xi)^2}
\ln\left(\sqrt{1+(\eta-\alpha+\xi)^2}-\eta+\alpha-\xi\right)\nonumber\\
&&-{1\over\alpha}\sqrt{1+(\eta-\alpha-\xi)^2}
\ln\left(\sqrt{1+(\eta-\alpha-\xi)^2}-\eta+\alpha+\xi\right)\nonumber\\
&&+{1\over4\alpha}
{2\alpha^2\beta^2(-3+\beta^2)+(-2+\beta^2)\beta^2+4\alpha(\beta^2-1)\xi
\over\sqrt{\beta^2+(\eta+\xi)^2}}
\ln{\sqrt{\beta^2+(\eta+\xi)^2}-\eta-\xi\over\beta}\nonumber\\
&&+{1\over4\alpha}
{2\alpha^2\beta^2(-3+\beta^2)+(-2+\beta^2)\beta^2-4\alpha(\beta^2-1)\xi
\over\sqrt{\beta^2+(\eta-\xi)^2}}
\ln{\sqrt{\beta^2+(\eta-\xi)^2}-\eta+\xi\over\beta}\nonumber\\
&&-{2\beta}\sqrt{{\beta^2\over4}-1}
\ln\left(\sqrt{{\beta^2\over4}-1}+{\beta\over2}\right)
\Biggr]\label{stamm1}
\end{eqnarray}
To simplify this expression we employ tho following substitution \cite{kamada}:
\begin{eqnarray}
\beta&\equiv&{u^2+1\over u}\label{defa1}\\
\alpha&\equiv&{2nm\over n^2-m^2}\label{defa2}
\end{eqnarray}
In terms of these new variables we get rid of all the square roots in
Eq. (\ref{stamm1}) and also the arguments of the logarithms simplify
greatly:
\begin{eqnarray}
\lim_{\Lambda\to\infty}{\cal I}_\alpha^\Lambda(\beta)=-2\pi\Biggl[
&&-{(m-n)^2+(m+n)^2u^4\over4mnu^2}\left(\ln{-m+n\over m+n}-2\ln u\right)
\nonumber\\
&&-{(m+n)^2+(m-n)^2u^4\over4mnu^2}\left(\ln{-m+n\over m+n}+2\ln u\right)
\nonumber\\
&&+{(m-n)^2+(m+n)^2u^4\over4mnu^2}\left(\ln{-m+n\over m+n}-\ln u\right)
\nonumber\\
&&+{(m+n)^2+(m-n)^2u^4\over4mnu^2}\left(\ln{-m+n\over m+n}+\ln u\right)
\nonumber\\
&&+{1-u^4\over u^2}\ln u\Biggr]
\label{stamm2}
\end{eqnarray}
One easily checks by comparing the coefficients of equal logarithms
that we get the final result
\begin{eqnarray}
\lim_{\Lambda\to\infty}{\cal I}_\alpha^\Lambda(\beta)&=&0\label{stamm3}\\
\beta&\ge&2\label{condition}
\end{eqnarray}
Finally, since $\lim_{\Lambda\to\infty}{\cal I}_\alpha^\Lambda(\beta)$
is an analytical function
which is zero over some finite interval on the real axis, we conclude
that
\begin{equation}
\lim_{\Lambda\to\infty}{\cal I}_\alpha^\Lambda(\beta)\equiv0\label{stamm4}
\end{equation}
everywhere on the complex $\beta$ plane and thus also for
$\beta=\beta_0$.
Because of Eq. (\ref{ibeta0})
we can then rewrite Eq. (\ref{split}) as
\begin{equation}
\int \hskip-4pt d^3 q\ F(\fett p,\fett q)=\int \hskip-4pt d^3 q\ F(0,\fett q)\label{pindep}
\end{equation}
demonstrating clearly the independence of $\fett p$.
As a consequence we can
give an expression for $m_2$ which is now, in contrast to
Eq. (\ref{m2}), independent of the initial momentum $\fett p$:
\begin{equation}
m_2=-{g^2\over8m_0(2\pi)^3}\int \hskip-4pt d^3 q\ {1\over E_{\Fett q}\omega_{\Fett
q}}\left({1\over\omega_{\Fett q}+m_0+E_{\Fett q}}+
{1\over\omega_{\Fett q}-m_0+E_{\Fett q}}\right)
\label{m2final}
\end{equation}
We arrive then indeed at Eq. (\ref{expect}):
\begin{equation}
\eta H'_{\mbox{\tiny nv}}=\int \hskip-4pt d^3 p\ \sqrt{m^2+\fett p^2}
a_{\Fett p}^\dagger|0\rangle\langle0|a_{\Fett p}+{\cal O}(g^3)
\label{expect1}
\end{equation}
with
\begin{equation}
m=m_0-{g^2\over8m_0(2\pi)^3}\int \hskip-4pt d^3 q\ {1\over E_{\Fett q}\omega_{\Fett
q}}\left({1\over\omega_{\Fett q}+m_0+E_{\Fett q}}+
{1\over\omega_{\Fett q}-m_0+E_{\Fett q}}\right)+{\cal O}(g^3)
\end{equation}
We want to point out that this result has been derived in the
framework
of a Hamilton formalism and did not use the connection to the standard
techniques.
|
\section{Introduction}
Since Stoler et al. introduced the binomial states (BSs)[1], the so-called
intermediate states have attracted considerable attention of physicists in
the field of quantum optics. A feature of these states is that they
interpolate between two fundamental quantum states, such as the number,
coherent and squeezed states, and reduce to them in two different limits.
For instance, the BSs interpolate between the coherent states (the most
classical) and the number states (the most nonclassical)[1-6], while the
negative binomial states (NBSs) interpolate between the coherent states and
geometric states[7-11]. Another feature of some intermediate states is that
their photon number distributions are some famous discrete probability
distributions in probability theory: the BS corresponds to the binomial
distribution, the NBS to the negative binomial distribution, the
hypergeometric state[12] to the hypergeometric distribution, and the
negative hypergeometric state[13] to the negative hypergeometric
distribution.
Recently Barnett introduced a new definition of NBS[14],
\begin{eqnarray}
|\eta ,M\rangle &=&\sum_{n=M}^\infty C_n(\eta ,M)|n\rangle \nonumber \\
&=&\sum_{n=M}^\infty \left[ {%
{{n} \choose {M}}%
}\eta ^{M+1}(1-\eta )^{n-M}\right] ^{1/2}|n\rangle ,
\end{eqnarray}
where $|n\rangle $ is the usual number state, $0<\eta \le 1$ and $M$ is a
non-negative integer. They find that the NBS $|\eta ,M\rangle $ and the BS
have similar properties if the roles of the creation operator $a^{\dagger }$
and annihilation operator $a$ are interchanged. The photon number
probability $|C_n^{}(\eta ,M)|^2$ is associated with the probability that $n$
photons were present given that $M$ are found and that the probability for
successfully detecting any single photon is $\eta $. Mixed states with the
photon number probability include those applicable to photodetection and
optical amplification[15].
The BS is a intermediate number-coherent state and the original NBS is a
intermediate geometric-coherent state. One question naturally arises that if
there exist an intermediate state which interpolates between the number and
geometric state. In reality, the new NBS is just the intermediate
number-geometric state. This fact will be seen in the next section. Thus, we
have three intermediate states which interpolate between two of the three
fundamental states (the number, coherent and geometric states).
In the present paper we shall study the nonclassical properties and
algebraic characteristics of the new NBS. The ladder operator formalism ,
displacement operator formalism and related algebraic structure will be
formulated in section II. It is interesting that the algebraic structure is
the SU(1,1) Lie algebra via the generalized Holstein-Primarkoff realization.
In section III, we will show that the NBS can be viewed as excited geometric
states, intermediate number-geometric states and nonlinear coherent states.
The nonclassical properties, such as sub-Poissonian distribution and
squeezing effect will be investigated in detail in section IV. The Q and
Wigner functions are studied in Section V and the two methods of generation
of the NBS are proposed in section VI. A conclusion is given in section VII.
\section{Ladder operator formalism ,displacement operator formalism and
algebraic structure of the new NBS}
\subsection{Ladder operator formalism and algebraic structure}
It is known that the BSs are special SU(2) coherent states[5,6] and the
original NBSs are Perelomov's SU(1,1) coherent states[11] via the standard
Holstein-Primakoff realizations. So we expect that the algebra involved in
the new NBS is SU(1,1) Lie algebra.
It is easy to evaluate that
\begin{equation}
a^{\dagger n}|\eta ,M\rangle =\left[ \frac{(M+n)!}{M!\eta ^n}\right]
^{1/2}|\eta ,M+n\rangle .
\end{equation}
In particular, for $n=1$, we get
\begin{equation}
a^{\dagger }|\eta ,M\rangle =\left( \frac{M+1}\eta \right) ^{1/2}|\eta
,M+1\rangle .
\end{equation}
The creation operator raises the NBS $|\eta ,M\rangle $ to $|\eta
,M+1\rangle $. This property is similar to the action of the creation
operator on the Fock state $|M\rangle $,
\begin{equation}
a^{\dagger }|M\rangle =\sqrt{M+1}|M+1\rangle .
\end{equation}
Actually, in the limit of $\eta \rightarrow 1$, the NBS $|\eta ,M\rangle
^{-} $ reduces to the number state $|M\rangle $ and Eq.(3) naturally reduces
to Eq.(4).
The key and interesting point is that there exists another operator $\sqrt{%
\hat{N}-M}$ which also raises the NBS $|\eta ,M\rangle $ to $|\eta
,M+1\rangle $. From Eq.(1), the following equation is directly derived as
\begin{equation}
\sqrt{\hat{N}-M}|\eta ,M\rangle =\left( {\frac{1-\eta }\eta }\right) ^{1/2}%
\sqrt{M+1}|\eta ,M+1\rangle .
\end{equation}
Comparing Eq.(3) and Eq.(5), we get
\begin{equation}
\sqrt{\hat{N}-M}|\eta ,M\rangle =\sqrt{1-\eta }a^{\dagger }|\eta ,M\rangle .
\end{equation}
Multiplying the both sides of the above equation by the operator $\sqrt{\hat{%
N}-M}$ from left, we obtain the ladder operator formalism of the NBS as
\begin{equation}
(\hat{N}-\sqrt{1-\eta }\sqrt{\hat{N}-M}a^{\dagger })|\eta ,M\rangle =M|\eta
,M\rangle .
\end{equation}
In the limit of $\eta \rightarrow 1$, we find that the NBS$|\eta ,M\rangle $
reduces to the number state $|M\rangle $ and Eq.(7) to the equation $\hat{N}%
|M\rangle =M|M\rangle $ as expected.
It can be proved that the operators appearing in Eq.(7) can form SU(1,1) Lie
algebra:
\begin{equation}
K_0=\hat{N}-\frac{M-1}2,K_{+}=\sqrt{\hat{N}-M}a^{\dagger },K_{-}=a\sqrt{\hat{%
N}-M}.
\end{equation}
Reminding of the standard Holstein-Primarkoff realization of SU(1,1) Lie
algebra, $J_0=\hat{N}+\frac M2,J_{+}=\sqrt{M+\hat{N}-1}a^{\dagger },J_{-}=a%
\sqrt{M+\hat{N}-1}$, we call the new realization as generalized
Holstein-Primakoff realization. To our knowledge, the new realization of
SU(1,1) Lie algebra seems not to be addressed in the literature.
In terms of the generators $K_0,K_{+}$ and $K_{-}$ of the SU(1,1) Lie
algebra , Eq.(7) is rewritten as
\begin{equation}
(K_0-\sqrt{1-\eta }K_{+})|\eta ,M\rangle =\frac{M+1}2|\eta ,M\rangle .
\end{equation}
\subsection{Displacement operator formalism}
Now we try to find the displacement operator formalism of the NBS. To this
end, let us rewrite the NBS as
\begin{eqnarray}
|\eta ,M\rangle &=&\eta ^{\frac{M+1}2}\sum_{n=0}^\infty \left[ {{%
{{M+n} \choose {M}}%
}}(1-\eta )^n\right] ^{1/2}|M+n\rangle \nonumber \\
&=&\eta ^{\frac{M+1}2}\sum_{n=0}^\infty \frac{(\sqrt{1-\eta }a^{\dagger })^n%
}{\sqrt{n!}}|M\rangle .
\end{eqnarray}
Then by making use of the following identity
\begin{equation}
\lbrack f(\hat{N})a^{\dagger }]^n=a^{\dagger n}f(\hat{N}+n)f(\hat{N}%
+n-1)...f(\hat{N}+1),
\end{equation}
Eq.(10) can be written in the exponential form
\begin{equation}
|\eta ,M\rangle =\eta ^{\frac{M+1}2}e^{{\sqrt{1-\eta }K_{+}}}|M\rangle .
\end{equation}
Note that $K_{-}|M\rangle =0$, then the displacement operator formalism is
obtained as
\begin{eqnarray}
|\eta ,M\rangle &=&e^{\sqrt{1-\eta }K_{+}}\eta ^{K_0}e^{-\sqrt{1-\eta }%
K_{-}}|M\rangle \nonumber \\
&=&e^{\xi (K_{+}-K_{-})}|M\rangle ,
\end{eqnarray}
where $\xi $=arctanh $\sqrt{1-\eta }$. In the derivation of the above
equation, we have used the identity
\begin{equation}
e^{\alpha K_{+}-\alpha ^{*}K_{-}}=e^{\gamma K_{+}}(1-|\gamma
|^2)^{K_0}e^{-\gamma ^{*}K_{-}},
\end{equation}
where $\gamma =\alpha \tanh |\alpha |/|\alpha |$. As seen from Eq.(13), the
NBS can be simply recognized as SU(1,1) displaced number states. Actually,
the NBS are essentially Peremolov's coherent states as shown below.
On the space
\begin{equation}
S=\text{span}\left\{ {|n+M\rangle \equiv |n;k\rangle |n=0,1,2...}\right\} ,k=%
\frac{M+1}2
\end{equation}
we have
\begin{eqnarray}
K_{+}|n;k\rangle &=&\sqrt{(n+1)(2k+n)}|n+1;k\rangle \nonumber \\
K_{-}|n;k\rangle &=&\sqrt{n(2k+n-1)}|n-1;k\rangle \nonumber \\
K_0|n;k\rangle &=&(n+k)|n;k\rangle
\end{eqnarray}
This is the discrete representation of SU(1,1) Lie algebra with Bargaman
index $k=(M+1)/2$. We see that the generalized Holstein-Primakoff
realization gives rise to the representation of SU(1,1) on the space $S$.
Note that $|M\rangle =|0;k\rangle $, the NBS can be written as
\begin{equation}
|\eta ,M\rangle =e^{\xi (K_{+}-K_{-})}|0;k\rangle .
\end{equation}
This shows the NBS are essentially Peremolov's coherent states.
\section{The NBS as excited geometric state, intermediate number-geometric
state and nonlinear coherent state}
\subsection{As excited geometric states and intermediate number-geometric
states}
From Eq.(10) we obtain
\begin{eqnarray}
|\eta ,M\rangle &=&\eta ^{(M+1)/2}\sum_{n=0}^\infty {%
{M+n \choose M}%
}^{1/2}(1-\eta )^{n/2}|M+n\rangle \nonumber \\
&=&\frac{\eta ^{(M+1)/2}}{\sqrt{M!}}a^{\dagger M}\sum_{n=0}^\infty (1-\eta
)^{n/2}|n\rangle \nonumber \\
&=&\frac{\eta ^{M/2}}{\sqrt{M!}}a^{\dagger M}|\eta \rangle _g,
\end{eqnarray}
where
\begin{equation}
|\eta \rangle _g=\eta ^{1/2}\sum_{n=0}^\infty (1-\eta )^{n/2}|n\rangle
\end{equation}
is the geometric state[16-21], which is also called Susskind-Glogower phase
state[11], phase eigenstate[16,17],and coherent phase state[19]. The photon
number distribution is $\eta (1-\eta )^n$, the geometric distribution. From
Eq.(18) the NBSs can be generated by repeated application of the creation
operator $a^{\dagger }$ on the geometric states. This shows that the NBS
belongs to an interesting class of nonclassical states, excited quantum
states. These states are first introduced by Agarwal and Tara as excited
coherent states[22]. So the NBSs can be viewed as excited geometric states.
Setting $M=0$ in Eq.(18), we get
\begin{equation}
|\eta,0\rangle=|\eta\rangle_g,
\end{equation}
which shows that the NBS reduces to the geometric states for $M=0$. Note
that the NBS reduces to the number state $|M\rangle$ in the limit of $%
\eta\rightarrow 1$. Thus, the NBS interpolates between the number state and
geometric state and can be viewed as number-geometric state.
\subsection{As nonlinear coherent states}
The geometric states are the eigenstates of the Susskind-Glogower phase
operator[23] $(1+N)^{1/2}a$, obeying the equation
\begin{equation}
(N+1)^{-1/2}a|\eta \rangle _g=\sqrt{1-\eta }|\eta \rangle _g.
\end{equation}
Comparing with the definition of the nonlinear coherent states$|\alpha
\rangle _{nl}$[24,25]
\begin{equation}
f(N)a|\alpha \rangle _{nl}=\alpha |\alpha \rangle _{nl},
\end{equation}
we know that the geometric states are nonlinear coherent states with the
nonlinear function $f(N)=(N+1)^{-1/2}$.
In a previous work, we have proved a general result that the excited
nonlinear coherent states are still nonlinear coherent states[26]. Since the
geometric states are nonlinear coherent states and the NBSs can be
recognized as excited geometric states, we infer that the NBSs are nonlinear
coherent states. In fact, multiplying Eq.(7) by the annihilation operator $a$
from the left , we get
\begin{equation}
(N+1-M)a|\eta ,M\rangle =\sqrt{1-\eta }\sqrt{N+1-M}(1+N)|\eta ,M\rangle .
\end{equation}
Since we discuss the problem in the space $S$(Eq.(15)), we can multiply
Eq.(23) by $1/[(1+N)\sqrt{N+1-M}]$ from left. This leads to
\begin{equation}
\lbrack \sqrt{N+1-M}/(N+1)]a|\eta ,M\rangle =\sqrt{1-\eta }|\eta ,M\rangle .
\end{equation}
The above equation shows that the NBS are nonlinear coherent states with the
nonlinear function $\sqrt{N+1-M}/(N+1)$. Eq.(24) naturally reduces to
Eq.(21) for $M=0$.
\section{Nonclassical properties}
\subsection{Sub-Poissonian distribution}
The simplest way to investigate the statistical characteristics of the
radiation field is to differentiate the generation function
\begin{equation}
G(\lambda )=\sum_{n=0}^\infty P(n)\lambda ^n=\lambda ^M\left( \frac \eta {%
1+\lambda \eta -\lambda }\right) ^{M+1}
\end{equation}
with respect to the auxiliary real number $\lambda $. Here $%
P(n)=|C_n^{}(\eta ,M)|^2$ is the photon distribution function of the NBS.
The factorial moments are defined as $F(n)=\frac{d^nG}{d\lambda ^n}%
|_{\lambda =1}$. From Eq.(25) we obtain the factorial moments $F(1)$ and $%
F(2)$ as
\begin{eqnarray}
F(1) &=&\langle N\rangle =\frac{M+1}\eta -1, \\
F(2) &=&\langle N^2\rangle -\langle N\rangle =\frac{(M+2)(M+1)}{\eta ^2}-4%
\frac{M+1}\eta +2.
\end{eqnarray}
Then we can easily derive Mandel's Q parameter
\begin{eqnarray}
Q &=&\frac{\langle N^2\rangle -\langle N\rangle ^2-\langle N\rangle }{%
\langle N\rangle } \nonumber \\
&=&\frac{F(2)-F^2(1)}{F(1)}=\frac{\eta ^2-2(M+1)\eta +M+1}{\eta (M+1-\eta )},
\end{eqnarray}
which measures the deviation from the Poisson distribution which corresponds
to the coherent state with $Q=0$. If $Q<0(>0)$, the field is called
sub(super)-Poissonian. The denominator of Eq.(28) is positive since $\eta
\le 1$, while the numerator can be positive or negative. For $M=0$, $%
Q=(1-\eta )/\eta \ge 0$. The NBS $|\eta ,0\rangle $(geometric state) is
super-Poissonian except $\eta =1$. For $M>0$, the condition for the
numerator $\eta ^2-2\eta (M+1)+M+1<0$ is
\begin{equation}
\eta >\eta _{-}=M+1-\sqrt{M(M+1)}.
\end{equation}
It can be proved that $0<\eta _{-}<1$. Thus the NBS $|\eta ,M\rangle
^{-}(M>0)$ is super-Poissonian when $\eta <\eta _{-}$ and sub-Poissonian
when $\eta >\eta _{-}$. As $M$ increases, $\eta _{-}$ decreases and the
sub-Poissonian range increases.
\subsection{Squeezing effect}
Define the quadrature operators $X$(coordinate) and $Y$ (momentum) by
\begin{equation}
X=\frac 12(a+a^{\dagger }),Y=\frac 1{2i}(a-a^{\dagger }).
\end{equation}
Then their variances
\begin{equation}
Var(X)=\langle X^2\rangle -\langle X\rangle ^2,\langle Var(Y)=\langle
Y^2\rangle -\langle Y\rangle ^2
\end{equation}
obey the Heisenberg's uncertainty relation
\begin{equation}
Var(X)Var(Y)\ge \frac 1{16}.
\end{equation}
If one of the $Var(X)$ and $Var(Y)$ is less than 1/4, the squeezing occurs.
In the present case, $\langle a\rangle $ and $\langle a^2\rangle $ are real.
Thus, the variances of $X$ and $Y$ can be written as
\begin{eqnarray}
Var(X) &=&\frac 14+\frac 12(\langle a^{\dagger }a\rangle +\langle a^2\rangle
-2\langle a\rangle ^2), \\
Var(Y) &=&\frac 14+\frac 12(\langle a^{\dagger }a\rangle -\langle a^2\rangle
).
\end{eqnarray}
From Eq.(2), the expectation values $\langle a\rangle $ and $\langle
a^2\rangle $ are obtained as
\begin{eqnarray}
\langle a\rangle &=&\eta ^{M+1}(1-\eta )^{-M}\sum_{n=M}^\infty {%
{n+1 \choose M}%
}^{1/2}{%
{n \choose M}%
}^{1/2}(1-\eta )^{n+1/2}(n+1)^{1/2} \\
\langle a^2\rangle &=&\eta ^{M+1}(1-\eta )^{-M}\sum_{n=M}^\infty {%
{n+2 \choose M}%
}^{1/2}{%
{n \choose M}%
}^{1/2}(1-\eta )^{n+1}[(n+2)(n+1)]^{1/2}
\end{eqnarray}
Using Eqs.(26), and (33)-(36), we can investigate the squeezing effect.
By numerical calculations, we find that the squeezing occurs in both the
quadrature $X$ and $Y$. Fig.1 gives the variance of the quadrature $X$
versus $\eta $ for different $M$. It can be seen that the range and degree
of squeezing increase as $M$ increases. There exists a critical value of $M$%
. When $M<7$, there is no squeezing for arbitrary values of $\eta $. The
squeezing also occurs in the quadrature $Y$ as shown in Fig.2. In contrary
to the squeezing in the quadrature $X$, the range and degree decrease as $M$
increases. When $M$ is larger than a critical value 31, no squeezing occurs.
\section{The Q and Wigner functions}
If a field is prepared in a quantum state described by a density operator $%
\rho $, we can define the $s$-parametrized quasiprobability distribution in
phase space as[27]
\begin{equation}
P(\beta ,s)=\frac 1{\pi ^2}\int d^2\xi C(\xi ;s)\exp (\beta \xi ^{*}-\beta
^{*}\xi ),
\end{equation}
where the quantum characteristic function is
\begin{equation}
C(\xi ,s)=Tr[D(\xi )\rho ]\exp (s|\xi |^2/2).
\end{equation}
Here $\beta =x+iy$, with $(x,y)$ being the $c$ numbers corresponding to the
quadratures $(X,Y)$, and $D(\xi )=\exp (\xi a^{\dagger }-\xi ^{*}a)$ is
Glauber's displacement operator. It is possible to write the $s$%
-parametrized quasiprobability distribution as a infinite series [28]
\begin{equation}
P(\beta ,s)=\frac 2\pi \sum_{n=0}^\infty (-1)^k\frac{(1+s)^k}{(1-s)^{k+1}}%
\langle \beta ,k|\rho |\beta ,k\rangle ,
\end{equation}
where $|\beta ,k\rangle =D(\beta )|k\rangle $ is the so-called displaced
number state. The expression above is suitable for direct numerical
calculations.
\subsection{Q function}
If we take $s=-1$, Eq.(39) is reduced to the familiar expression for the $Q$
function
\begin{equation}
Q(\beta )=\frac 1\pi \langle \beta |\rho |\beta \rangle .
\end{equation}
Note that since the NBS can be viewed as an excited geometric state [see
Eq.(18)], we obtain the Q function of the NBS as
\begin{equation}
Q(\beta )=\eta ^{M+1}\exp (-|\beta |^2)|\beta |^{2M}\left| \sum_{n=0}^\infty
\beta ^n(1-\eta )^{n/2}/\sqrt{n!}\right| ^2/M!.
\end{equation}
In Fig.3 we present plots of the $Q$ function of a NBS for different values
of $\eta $ and $M=5$. We can clearly see the deformation of the $Q$
function. When $\eta =1$, the $Q$ function representing the number state $%
|M=5\rangle $ is formed[Fig3.(d)] as expected. From the $Q$ function we can
also study the squeezing effects by examining the deformation of their
contours. Fig.4 is the contour plot of the $Q$ functions of two particular
NBSs, with (a) $\eta =0.2,M=5$ and (b) $\eta =0.85,M=50$. We can clearly see
the compression along the $y$ and $x$ direction, which corresponds to
squeezing in the $Y$ quadrature of the first NBS and in the $X$ quadrature
of the second.
\subsection{Wigner function}
By taking $s=0$ in Eq.(39), we obtain a series representation for the Wigner
function
\begin{equation}
W(\beta )=\frac 2\pi \sum_{k=0}^\infty (-1)^k\langle \beta ,k|\rho |\beta
,k\rangle .
\end{equation}
Now we insert Eq.(1) into the above equation, which yields
\begin{equation}
W(\beta )=\frac 2\pi \sum_{k=0}^\infty \left| \sum_{n=M}^\infty C_n^{}(\eta
,M)\chi _{nk}(\beta )\right| ^2.
\end{equation}
In the expression above, the matrix elements $\chi _{nk}(\beta )=\langle
n|D(\beta )|k\rangle $ are given by[29]
\begin{equation}
\chi _{nk}(\beta )=\beta ^n(-\beta ^{*})^k\exp (-|\beta
|^2/2)_2F_0(-n,-k;|\beta |^{-2})/\sqrt{n!k!},
\end{equation}
where $_2F_0(\alpha ,\beta ;z)$ are the generalized hypergeometric
functions[30]. The present form of $\chi _{nk}(\beta )$ is convenient for
numerical calculations.
The Wigner function can be used to trace the nonclassical behaviors of
quantum states. It is known that the negativity of the Wigner function is a
sufficient but not necessary condition for having nonclassical effects. In
Fig.5 we give plots of the Wigner function of a NBS by numerical
calculations of Eq.(43) for different values of $\eta$ and $M=1$. As in
Fig.5(a), the negative part of the Wigner function is already noticeable for
$\eta=0.3$. For $\eta=0.5$[Fig.5(b)], the negative part is pronounced. In
fig5(c), for $\eta=0.9$, the negative part is even larger, and finally, in
Fig.7(d), we have the full Wigner function of a number state $|1\rangle$($%
\eta=1$). The Wigner function becomes more and more negative as $\eta$
increases.
\section{Generation of the new NBS}
Let us discuss the dynamical generation of the NBS. We consider two
different methods. The first is quite straightforward in concept from the
displacement operator formalism(Eq.(13)) but might not be very easy to
achieve experimentally. The Hamiltonian is given by
\begin{eqnarray}
H &=&H_0+i\chi [\sqrt{\hat{N}-M}a^{\dagger }\exp (-i\omega t)-a\sqrt{\hat{N}%
-M}\exp (i\omega t)] \nonumber \\
H_0 &=&\omega a^{\dagger }a.
\end{eqnarray}
The constant $\chi $ is the coupling strength. The coupling is of
intensity-dependent type and is similar to those in some
intensity-dependent Jaynes-Cummings models [31-34]. The unitary time
evolution operator in the interaction picture is
\begin{equation}
U(t)=\exp [\chi t(\sqrt{\hat{N}-M}a^{\dagger }-a\sqrt{\hat{N}-M})].
\end{equation}
Supposing the system is initially prepared in the number state $|M\rangle $,
we find the system at time $t$ is the NBS
\begin{equation}
U(t)|M\rangle =|1-\tanh ^2(\chi t),M\rangle .
\end{equation}
The second method of the generation of the NBS is based on the fact that the
NBS is the excited geometric state. The geometric state can be prepared in
the non-degenerate three-wave interaction system[35]. We can also generate
the geometric state by the non-degenerate parametric amplifier described by
the two-mode Hamiltonian[36]
\begin{eqnarray}
\tilde{H} &=&\tilde{H}_0+i\chi [a_1^{\dagger }a_2^{\dagger }\exp (-2i\omega
t)-a_1a_2\exp (2i\omega t)], \nonumber \\
\tilde{H}_0 &=&\omega _1a_1^{\dagger }a_1+\omega _2a_2^{\dagger }a_2,
\end{eqnarray}
where $a_1$ and $\omega _1$ ($a_2$ and $\omega _2$) are the annihilation
operator and frequency for the signal(idler) mode. Frequencies $\omega _1$
and $\omega _2$ sum to the pump frequency, $2\omega =\omega _1+\omega _2$.
The coupling constant $\chi $ is proportional to the second-order
susceptibility of the medium and to the amplitude of the pump. The unitary
time evolution operator in the interaction picture is
\begin{equation}
\tilde{U}(t)=\exp [\chi t(a_1^{\dagger }a_2^{\dagger }-a_1a_2)]
\end{equation}
Suppose that the system is initially prepared in the state $|0,0\rangle
=|0\rangle _1\otimes |0\rangle _2.$ Then at any time $t$ the system is in
the state
\begin{equation}
|\eta \rangle _{tm}=\tilde{U}(t)|0,0\rangle =\eta ^{1/2}\sum_{n=0}^\infty
(1-\eta )^{n/2}|n,n\rangle
\end{equation}
which is the two-mode geometric state in comparison with Eq.(19). Here $\eta
=1-\tanh ^2(\chi t).$
Once the two-mode geometric state is prepared, one can generate two-mode
NBS by the following procedure in analogy to that proposed by Agawal and
Tara[22]. Consider the passage of a two-level excited atom through a cavity.
Let the initial state of the atom-field system be $|\eta \rangle
_{tm}\otimes |e\rangle ,$ where $|e\rangle $ is the atomic excited state.
The interaction Hamiltonian has the form[37]
\begin{equation}
\bar{H}=\hbar (gS^{+}a_1+g^{*}S^{-}a_1^{\dagger }),
\end{equation}
where $S^{\pm }$ are the psedospin operators of the atom and $g$ is the
coupling constant. Since $g$ is generally small, the state at time $t$ can
be approximated by
\begin{equation}
|\psi (t)\rangle \approx |\eta \rangle _{tm}\otimes |e\rangle
-ig^{*}ta_1^{\dagger }|\eta \rangle _{tm}\otimes |g\rangle ,
\end{equation}
which is valid for interaction times $gt<<1.$ From the above equation we
observe that, if the atom is detected to be in the ground state $|g\rangle ,$
then the state of the field is reduced to $a_1^{\dagger }|\eta \rangle _{tm}.
$ An extension of the above arguments to the multiphoton processes would
imply that the state $a_1^{\dagger M}|\eta \rangle _{tm}$ can be produced in
multiphoton processes. For the multiphoton processes, the Hamiltonian
(Eq.(51)) is replaced by a new Hamiltonian with $a_1\rightarrow a_1^M.$
Thus, the above procedure for a multiphoton process will result in the state
\begin{equation}
|\eta ,M\rangle _{tm}=\eta ^{(M+1)/2}\sum_{n=0}^\infty {%
{M+n \choose M}%
}^{1/2}(1-\eta )^{n/2}|M+n,n\rangle .
\end{equation}
The above normalized state is just the two-mode NBS. In a short summary
two
methods are proposed to generate the NBS.
\section{Conclusions}
We have investigated the NBS induced recently by Barnett and found the
ladder operator formalism and displacement operator formalism of the NBS.
The algebra involved is SU(1,1) Lie algebra via the generalized
Holstein-Primakoff realization. We found that the NBS are essentially
Peremolov's coherent states.
As excited quantum states, the NBSs are excited geometric states. As
intermediate states, they interpolate between the number and geometric
states. According to the definition of the nonlinear coherent states, we
find that the NBSs are nonlinear coherent states.
The NBS can be sub-Poissonian or super-Poissonian. There exists a critical
point $\eta _{-}$. The NBS is sub-Poissonian when $\eta >\eta _{-}$ and
super-Poissonian when $\eta <\eta _{-}$. The squeezing occurs in both the
quadrature $X$ and $Y$ with two critical values of $M=7$ and 31,
respectively. There is no squeezing occurs in the quadrature $X$ for $M<7$
and in the quadrature $Y$ for $M>31$ for arbitrary values of $\eta $. The $Q$
and Wigner functions are studied numerically. They show that the NBSs have
prominent nonclassical properties. We have proposed two methods of generation
of the new NBS.
In addition, the remarkable properties of the new NBS seem to suggest that
it deserves further attention from both theoretical and application sides of
quantum optics.
\vspace{1cm}
{\bf Acknowledgment}: One of the authors(Wang) thanks for the discussions
with Prof. H.C.Fu and help of Prof. C.P.Sun. The work is partially supported
by the National Science Foundation of China with grant number:19875008.
\newpage
|
\section{introduction}
\label{intro}
Ferrofluids~\cite{rr} are oil- or water-based colloidal suspensions of
permanently magnetized particles. In an applied magnetic field the particles
align creating a strong paramagnetic response in the ferrofluid. Because they
are fluids, these suspensions can flow in response to forces. For example,
ferrofluid droplets elongate parallel to applied
fields~\cite{d3droplets1,d3droplets2,d3droplets3,ellipsoid,bashtovoi,d2droplets}
and undergo tip-sharpening
transitions~\cite{sharpen1,sharpen2}. When a ferrofluid droplet is confined
between two plates in a ``thin film'' geometry, surrounded by an
immiscible fluid, and a field is applied perpendicular to the plates,
it undergoes field induced bifurcations~\cite{goldstein} leading to
intricate labyrinthine patterns~\cite{pattern}. Ferrofluid
emulsions~\cite{bibette} undergo structural
transitions under an applied field from a randomly dispersed structure of
the emulsion droplets to droplet chains, columns and worm like
structures~\cite{liu1,liu2} depending on volume fraction, sample geometry
and the rate of field application.
A droplet of ferrofluid elongates under applied field because of
the demagnetizing fields of magnetic
poles on the surface of the droplet. Surface poles arise wherever the
droplet magnetization has a component perpendicular
to the surface. The demagnetizing field that they create opposes the
magnetization,
creating a demagnetizing energy that depends on the shape of the
droplet. The droplet elongates to reduce its demagnetizing field and
energy. Because elongation increases the surface energy of the
system, an equilibrium shape is reached when the magnetic forces balance
against the surface tension forces.
The elongation of freely suspended, 3-dimensional droplets
has been well studied~\cite{d3droplets1,d3droplets2,d3droplets3,ellipsoid}.
The droplets can be assumed to be
ellipsoids for small elongation. The demagnetizing field is thus uniform
and the elongation (major axis minus minor axis divided by minor axis) is
found to be proportional to the undeformed
droplet radius. The case of droplets confined in ``thin film geometry''
however, involves two length scales, droplet thickness and its undeformed
diameter. In the limit of small aspect ratio (droplet thickness divided by its
undeformed diameter) the demagnetizing fields are stronger near the edges
of the droplet than at its center. We find that the elongation
divided by droplet thickness in this geometry is proportional to
the logarithm of the aspect ratio.
Prior experiments~\cite{d2droplets,bashtovoi}
have proposed droplet elongation as a tool for measuring surface tension
between the ferrofluid and the surrounding immiscible fluid.
We improve on the existing theory~\cite{d2droplets} by incorporating
spatial variation of the demagnetizing field inside the droplet. We perform
an experiment supporting our predicted logarithmic behavior.
Section~\ref{theory} of this paper presents our theoretical study of the
elongation of a ferrofluid droplet confined within a thin film. Our principal
result is a predicted logarithmic dependence of elongation on droplet
aspect ratio. We contrast this result with the corresponding elongation of
unconfined droplets. Section~\ref{expt} describes an experiment done with
ferrofluid emulsions that tests our theory. The experiment is in
qualitative agreement with our theoretical prediction, but differs
quantitatively in at least one respect. In section~\ref{discussion} we discuss
a possible explanation of the discrepancy based upon droplet contact angles
with the confining plates.
\section{theory}
\label{theory}
Consider a paramagnetic liquid droplet confined in a thin film between two
parallel plates with a gap, $\Delta$, in the ${\hat z}$ direction
(see figure 1). An immiscible liquid surrounds the droplet. Let the thickness,
$\Delta$, be much smaller than the radius of the undeformed droplet, $r_0$.
This small aspect ratio
\begin{equation}
p={\Delta \over 2r_0}
\end{equation}
provides the pseudo-two-dimensional character of the problem. If a uniform,
weak, field ${\bf H}_0$ is applied parallel to the plate, the droplet
magnetizes. The magnetization creates an opposing demagnetizing
field whose strength depends on the droplet shape. The droplet elongates
to decrease its magnetic energy, reaching equilibrium when the magnetic
forces balance against the restoring forces due to surface tension.
In this section we define the elongation of the droplet and calculate
the surface energy, $E_S$, and the magnetic energy, $E_M$, of the droplet as
a function of its elongation. By minimizing the total energy with respect
to the elongation we obtain the elongation as a function of ${\bf H}_0$,
$r_0$, and $\Delta$.
For simplicity assume the elongated droplet has a uniform cross
section, $\cal C$, independent of $z$. This corresponds to a contact
angle of $90^\circ$ between the paramagnetic liquid, the surrounding fluid
and the glass plates, and a plate spacing much less than the capillary length
of the two liquids. Thus the droplet has straight edges if viewed from the
side (see figure 1). The role of
contact angle will be discussed later in section~\ref{discussion}.
We write the equation for $\cal C$ in polar coordinates as a generic
smooth perturbation to a circle,
\begin{equation}
\label{curve}
r=\alpha_1+\alpha_2 \cos 2 \theta.
\end{equation}
We only include a single harmonic, since we expect coefficients for
the higher harmonics to be much smaller than $\alpha_2$ for small
perturbations. The cross section $\cal C$ has semi-major axis $a$, and
semi-minor axis $b$ (see figure 1b), with $\alpha_1=(a+b)/2$ and
$\alpha_2=(a-b)/2$. We define the elongation of the droplet
\begin{equation}
\label{definition}
\epsilon \equiv {a \over b}-1.
\end{equation}
We assume that the elongation, $\epsilon$, is much less than $1$.
Imposing the constraint that the volume of the droplet ($\Delta$ times
cross-sectional area) remains constant we calculate
\begin{equation}
\label{alpha}
\alpha_1= {r_0 \over (1+k^2/2)^{1/2}},~~~~\alpha_2={r_0 k
\over (1+k^2/2)^{1/2}},
\end{equation}
where $k=\epsilon / (2+\epsilon)$.
The surface energy is the sum of interfacial areas times surface tensions
between all pairs of the three phases (solid glass, ferrofluid droplet and
immiscible fluid). For the case of uniform cross-section ($90^\circ$
contact angle) droplets, the glass-ferrofluid and glass-immiscible fluid
interfacial areas are independent of the shape of ${\cal C}$ due to the
fixed volume constraint. Hence we concern ourselves with the
droplet-surfactant solution interface, the area of which is $\Delta$ times
the perimeter. The perimeter of cross section $\cal C$ can be calculated
as a power series in $\epsilon$,
\begin{equation}
\label{circum}
S= 2 \pi r_0 (1+{3 \over 16} \epsilon^2+O(\epsilon^3)).
\end{equation}
As expected, the leading correction to $S$ is second order in $\epsilon$
since the perimeter should increase regardless of the sign of $\epsilon$.
The relevant surface energy of the droplet is
\begin{equation}
\label{surface-nrg}
E_S=\sigma_{FI} S \Delta
\end{equation}
where $\sigma_{FI}$ is the surface tension of the
ferrofluid-immiscible fluid interface.
The total magnetic energy of any paramagnetic body under applied field
is~\cite{wfb}
\begin{equation}
\label{magnetic-nrg}
E_M= -{1 \over 2}\int_V d^3{\bf r}~{\bf H}_0 \cdot {\bf M}({\bf r}).
\end{equation}
The magnetization ${\bf M}({\bf r})$ is determined by the self consistent
equation
\begin{equation}
\label{M}
{\bf M}({\bf r})=\chi({\bf H}_0+{\bf H}_D({\bf r}))
\end{equation}
for linear susceptibility $\chi$, where
\begin{equation}
\label{demag-field}
{\bf H}_D({\bf r})= \int_{S} d^2 {\bf r'}~
({\bf M}({\bf r'}) \cdot {\bf \hat{n}}({\bf r'}))
{ {\bf r-r'} \over |{\bf r-r'}|^3}
+\int_{V} d^3 {\bf r'}~
(\nabla \cdot{\bf M}({\bf r'}))
{{\bf r-r'} \over |{\bf r-r'}|^3}
\end{equation}
is the demagnetizing field due to the magnetization ${\bf M}({\bf
r})$, with ${\bf {\hat n}}({\bf r'})$ being the outward normal at any
point on the surface. The surface integral gives the demagnetizing
field due to the surface poles which appear wherever the magnetization
has a component normal to the surface. The volume integral gives the
contribution to the demagnetizing field due to volume charges which
appear at points where the magnetization has non-zero divergence.
To calculate the magnetic energy we expand ${\bf M}$ and ${\bf H}_D$
in power series in the susceptibility $\chi$,
\begin{equation}
\label{mag}
{\bf M}({\bf r})={\bf M}^{(1)}({\bf r})+{\bf M}^{(2)}({\bf r})
+{\bf M}^{(3)}({\bf r})+...
\end{equation}
\begin{equation}
\label{demag}
{\bf H}_D({\bf r})={\bf H}_D^{(1)}({\bf r})+{\bf H}_D^{(2)}({\bf r})+
{\bf H}_D^{(3)}({\bf r})+... ,
\end{equation}
where ${\bf M}^{(n)}({\bf r})$ and ${\bf H}_D^{(n)}({\bf r})$ are proportional
to $\chi^n$. Equating terms in~(\ref{M}) of equal order in $\chi$ we get
\begin{equation}
\label{first-order}
{\bf M}^{(1)}({\bf r})=\chi {\bf H}_0
\end{equation}
and
\begin{equation}
\label{higher-order}
{\bf M}^{(n+1)}({\bf r})=\chi {\bf H}_D^{(n)}({\bf r}).
\end{equation}
Note that ${\bf M}^{(1)}({\bf r})$ is independent of ${\bf r}$ because
the applied field is uniform whereas ${\bf M}^{(n)}({\bf r})$ may depend on
$({\bf r})$ for $n>1$ because ${\bf H}_D({\bf r})$ may be non-uniform. To
second order in $\chi$ we write the magnetic energy of the droplet
in~(\ref{magnetic-nrg}) as
\begin{equation}
\label{mag-nrg2}
E_M=-{1 \over 2}\chi{H}_0^2 V -{1 \over 2}\int_V d^3{\bf r}
~{\bf M}^{(1)}\cdot {\bf H}_D^{(1)}.
\end{equation}
The first term in equation~(\ref{mag-nrg2}) for the magnetic energy is
independent of the shape of the droplet and hence unimportant for our
consideration. The second term in the energy is the demagnetizing
energy $E_D$ due to a uniform magnetization ${\bf M}^{(1)}=\chi{\bf
H}_0$. Because ${\bf M}^{(1)}$ is uniform there are no volume charges,
and the surface poles appear only along the droplet-immiscible
fluid interface, to first order in $\chi$. Rewrite the second term
in~(\ref{mag-nrg2}) as an energy due to the induced surface charges
along the curved surface of the droplet
\begin{equation}
\label{demag-nrg}
E_D={1 \over 2} \chi^2 \int_0^\Delta dz \int_0^\Delta dz'
\oint ds \oint ds' {({\hat {\bf n}} \cdot {\bf H}_0)
({\hat {\bf n'}} \cdot {\bf H}_0) \over |{\bf r}-{\bf r'}|}.
\end{equation}
Here $ds$ and $ds'$ are infinitesimal arc-lengths along the contour of the
droplet ${\cal C}$, and ${\hat {\bf n}}$ and ${\hat {\bf n'}}$ are the outward
normals to the curved surface of the droplet at points $(s,z)$ and $(s',z')$
respectively.
Write $|{\bf r}-{\bf r'}|= {\sqrt {R^2+(z-z')^2}}$, where
$R$ is the in-plane distance between points at
positions $s$ and $s'$ on $\cal C$. Integrating over $z$ and $z'$
in~(\ref{demag-nrg}) yields~\cite{goldstein}
\begin{equation}
\label{demag-nrg2}
E_D= \chi^2 \Delta \oint ds \oint ds' ({\hat {\bf n}} \cdot {\bf H}_0)
({\hat {\bf n}}' \cdot {\bf H}_0) \Phi(R/ \Delta)
\end{equation}
where
\begin{equation}
\Phi (R/\Delta)= R/\Delta - {\sqrt {1+(R/\Delta)^2}} +
\ln {\Big [}(R/\Delta)/({\sqrt {1+(R/\Delta)^2}}-1){\Big ]}.
\end{equation}
Using equation~(\ref{curve}) for $\cal C$ we calculate the
demagnetizing energy in~(\ref{demag-nrg2}) as a series expansion in
$\epsilon$ and the aspect ratio $p=\Delta/2r_0$
\begin{equation}
\label{expansion}
E_D= \chi^2 H_0^2 V {\Big\{}{2p} \ln {B \over p}- 3\epsilon p
\ln { C \over p}+ \cdot \cdot \cdot {\Big \}}
\end{equation}
where $V=\pi r_0^2 \Delta$ is the volume of the droplet, and $B=4
e^{-1/2}$ and $C=4 e^{-5/6}$ are geometrical constants. The term in
the brackets can be identified as $2 \pi$ times the demagnetizing
factor~\cite{wfb}
of the droplet along the direction of applied field. Additional terms
in the series in equation~(\ref{expansion}) are of higher order in
$\epsilon$ or in $p$. For small elongation and large aspect ratio
we may neglect these higher order terms.
Minimizing the total energy $E=E_S+E_M$ with respect to $\epsilon$ gives
\begin{equation}
\label{elongation}
\epsilon= {\chi^2 H_0^2 \Delta \over \sigma_{FI}}\ln {C \over p}.
\end{equation}
Corrections to this result are higher order in aspect ratio $p$ or higher
order in $\epsilon$ itself. Interestingly, the elongation depends only
logarithmically on the undeformed radius $r_0$, and has a much stronger
dependence on the thickness, $\Delta$, of the droplet. This result differs
from an earlier theory~\cite{d2droplets} which omits the logarithm because
it assumes that the demagnetizing field is uniform inside the droplet.
In the case of unconfined, nearly ellipsoidal
droplets~\cite{d3droplets3,ellipsoid},
the demagnetizing field is quite uniform inside the droplet. The demagnetizing
energy is therefore proportional to the volume ($(4/3)\pi r_0^3$) of the
droplet according to equation~(\ref{mag-nrg2}). The surface energy is
proportional to the area ($4 \pi r_0^2$) and the elongation is thus
proportional to $r_0$. In the case of thin film geometry, however, the
demagnetizing field is very non-uniform. For distances much less than $\Delta$
near the droplet edge, the component of the demagnetizing field is of order
$M$, since the edge acts like an infinite sheet of charge in the first
approximation. For distances much greater than $\Delta$ the demagnetizing
field is of order $M \Delta/r$ since the edge acts as a line charge in
this case. The contribution to the integral for the
demagnetizing energy in equation~(\ref{mag-nrg2}) mainly comes from the bulk
of the droplet and goes like $r_0 \Delta^2 \ln (r_0/\Delta)$. The
surface energy is proportional to $2 \pi r_0 \Delta$ and the
elongation is therefore proportional to $\Delta \ln (r_0/\Delta)$. The
logarithmic variation of elongation with the aspect ratio is thus a
signature of the non-uniform nature of the demagnetizing field inside the
droplet.
\section{experiment}
\label{expt}
\subsection{Setup}
\paragraph{Sample Preparation and Structure}
Our sample consisted of a ferrofluid/aqueous solution emulsion confined
between two glass plates. The oil-based ferrofluid used was EMG 905 made
by Ferrofluidics. To reduce the surface tension between the ferrofluid and
the immiscible aqueous external phase, we incorporated surfactants in
the aqueous phase. A solution of a commercial detergent made the best
emulsions while solutions
with other pure anionic surfactants either showed hardly any elongation
of the ferrofluid droplets under applied field or produced droplets
without sharp boundaries with the aqueous phase. In contrast, our stable, well
behaved emulsions allowed us to probe and confirm the fundamental aspects
of our model.
To prepare the emulsions, a single drop of ferrofluid ($\sim 0.1$ ml)
was added to $10$ ml of surfactant solution which was a $12$ times dilution
of the commercial detergent. The liquid was shaken (by hand) to prepare the
emulsion, creating ferrofluid droplets with diameters varying from
$\sim 5-200~\mu$m. A small amount of this emulsion was then put between
two glass plates which were circular, about $2$ cm in diameter and $4$ mm
in thickness. These plates were cleaned using soap and alcohol and then rinsed
with ROPure water. We also tried acid cleaning of the glass plates, however it
did not result in any noticeable change in the quality of the sample.
We used a rectangular spacer made of mylar foil to separate the plates and
prevent the emulsion from leaking out from the edges of the plates. The mylar
foil extended to to the edges of the glass plates and had a rectangular
hole in the center into
which the emulsion was inserted. The thickness of a single mylar spacer was
measured to be $6.54\pm 0.06~\mu$m. The experiment was performed with one
and two spacers to ensure small aspect ratio.
For the cell assembly, the mylar spacers were placed on the first plate
and a drop of the emulsion was put in the center of the plate. The second
plate was placed on top and the two plates were clamped together using
a pair of brass rings. The rings were tightened by a set of $4$ equally
spaced screws. We measured the thickness variation across the sample by
making a ``dry'' sample (without the emulsion) and counting resulting
white light interference fringes. Although the thickness of mylar spacers
was measured to an accuracy of $1$ percent, the thickness variation across
the sample was found to be $10 \%$ resulting from the stresses
due to clamping and possible entrapment of dust in the cell.
\paragraph{Apparatus}
A schematic diagram of the experimental setup is shown in figure 2.
We put the sample at the center of a pair of Helmholtz coils
to insure a homogeneous magnetic field. The field measured
close to the sample using a Hall probe showed a variation of less than
$4\%$ across the sample. The sample was set up horizontally to prevent
gravitational settling of the ferrofluid droplets. Horizontal alignment
was achieved using a bubble level.
The sample was illuminated from below using a diffused light source and
observed from above using a tele-microscope. The
tele-microscope was connected to a CCD camera and the image from it was
fed into a video recorder and recorded on video tape. Images
from the recording were later processed using {\it NIH Image}. We calibrated
the optical system using a measuring reticule aligned along the two
orthogonal directions of the CCD array. Figure 3
shows a low magnification view of a typical sample. The ferrofluid droplets
appear much darker in the image than the surfactant solution around them.
\paragraph{Experimental Procedure and Image Analysis}
During the experiment the applied field was incremented every few seconds.
We found the response of the droplets to the field to be nearly
instantaneous and the shape of the droplets remained constant at constant
field. Experiments with decreasing field strength showed no hysteresis in
droplet shape. While droplet elongations were observed to be small we
incremented the field in steps of about $1$ Gauss, and increased the increments
up to about $5$ Gauss as the elongation increased. Droplet elongations appeared
to vary smoothly with applied fields over the entire range from $0$ to $50$
Gauss.
During each experiment the droplets were observed on a video monitor and
recorded on tape. After grabbing images of distorted droplets, we used a
cut-off in pixel gray scale level to identify the droplet edge. The
semi-major axis $(a)$ and the semi-minor axis $(b)$ were directly read off
the image using {\it NIH Image}. At zero field measured elongations were small
(RMS magnitude around $0.003$) and in random directions. These minor
perturbations from a circular shape were likely due to microscopic distortion
of the contact line pinned on weak surface heterogeneities. The ``observed
radius'' $r_0$ was calculated as the average of the two semi-axes at zero
field and the elongation at each field value was calculated using data
analysis software.
\paragraph{Results}
For each of the $48$ droplets studied we plotted
elongation $\epsilon$, versus the square of the applied field, ${\bf H}_0$.
Figure $4$ shows typical plots. The elongation is proportional to the
square of the applied field for small applied fields as predicted.
Saturation effects, although small, can be seen at higher values of the field.
The plot of elongation, for each droplet was fitted to
\begin{equation}
\label{fit}
{\epsilon \over \Delta}=k_0+k_1 H_0^2 + k_2 H_0^4.
\end{equation}
We included terms only up to order $H_0^4$ because the saturation effects
were observed to be small. We include $k_0$ to allow for the observed small
elongations at zero field.
The coefficients $k_1$ of each droplet were then plotted versus the inverse
of the aspect ratio $1/p=2 r_0 /\Delta$
on a semi-log plot (see figure 5). The theory predicts a slope of $\chi^2 /
\sigma_{FI}$ and an intercept of $1/C$ on the horizontal axis with $C=1.74$.
The data points in figure 5(a) fall on a straight line as predicted
by the theory. Also, as predicted by the theory, the data points for
two different droplet thicknesses overlay each other.
There is substantial scatter in the data, but the deviations from a
straight line are random and consistent with the error bars. The chief
source of uncertainty was the $10 \%$ uncertainty in thickness due to the
variation observed across the sample. Figure 5(b) displays the deviation
of $k_1$ from the best fit normalized by the uncertainty. The uncertainties
in measuring $\epsilon, r_0$ and ${\bf H}_0$ were found to be negligible in
comparison.
Dividing the susceptibility $\chi=1.9$ for the ferrofluid
used~\cite{ferrofluidics} by the slope
= $0.119\pm0.004$~cm/dyne obtained from the fitted line we get
$\sigma_{FI}=30.4 \pm 1.1$ dynes/cm,
typical of oil-water surface tensions.
From the fitted line we also get $C=0.35\pm 0.08$, differing substantially
from our theoretically predicted value of $1.74$. It may be
possible to explain this discrepancy by considering the effect of the
contact angle of the ferrofluid-immiscible fluid interface with the glass
plates. In the discussion section below we explore the qualitative
effect of the contact angle.
In figure 6(a) we plot $\epsilon /\Delta$ versus $2r_0/\Delta$ on a linear
scale. If the demagnetizing field inside the droplet was uniform like in
the case of unconfined droplets,
the plot would be a straight line. However, the plot is clearly not a straight
line and the deviations from the best fitted straight line are systematic (see
figure 6(b)).
This further supports our theoretical result that the demagnetizing field
inside a confined droplet is non-uniform and the elongation divided by
thickness is proportional to the logarithm of the aspect ratio.
\section{Discussion}
\label{discussion}
The results discussed in section~\ref{expt} agree with
our theoretical prediction~(\ref{elongation}) of logarithmic variation
of $\epsilon/\Delta$ with a proportionality constant of $\chi^2/\sigma_{FI}$.
However, our theoretical value for $C$ is $4e^{-5/6}=1.74$ whereas the
experimentally measured value for $C$ is $0.35 \pm 0.08$.
One possible explanation for the discrepancy in the value of $C$ is
that the ferrofluid/glass contact angle is not $90^\circ$ and consequently
the cross-section of the droplet is not uniform. Our calculations
are for uniform droplet cross-section, which corresponds to a contact
angle $\beta=90^\circ$ between the glass plate and liquid droplet.
The experiment, however, was performed with an oil-based ferrofluid in
a surfactant solution for which the oil-glass contact angle $\beta <
90^\circ$ (see figure 7). A contact angle of other than $90^\circ$
will affect the elongation in two ways: by changing interfacial areas
to alter the functional form of $E_S$ and by redistributing the magnetic
surface poles to alter the functional form of $E_M$. We consider these
two effects in turn. First, however, we must address an ambiguity in
the definition of aspect ratio and elongation which results from the
non-uniformity of droplet cross-section.
Our experiment observes the profile of the largest cross-section of
the droplet. For a circular droplet with $\beta<90^\circ$ this is the
radius $r_1$ defined as the radius at mid-gap as shown in figure 7.
For an elongated droplet we measure the semi-major and -minor
axes $a_1$ and $b_1$ and, through equation~(\ref{definition}), the
elongation $\epsilon_1$. We also define $r_2$, $a_2$, $b_2$ and
$\epsilon_2$ associated with the ferrofluid-immiscible fluid-glass
plate contact line (see figure 7). Since $\Delta$ is much less than the
capillary length of the ferrofluid/immiscible fluid, to a good
approximation~\cite{contact} the profile of
the droplet will be an arc of a circle, so the difference between
$r_1$ and $r_2$ is of order $\Delta$, and likewise for the semi-major
and -minor axes. The difference $\epsilon_1-\epsilon_2$ is of order
$\Delta/r_1$ relative to the elongation. Recall that our
result~(\ref{elongation}) for the elongation is only the lowest order
term in a series expansion in the aspect ratio. Thus the distinction
between $r_1$ and $r_2$, and between $\epsilon_1$ and $\epsilon_2$,
does not alter our result at the lowest order in aspect ratio.
When the contact angle differs from $90^\circ$, the cross section of
the droplet depends on $z$. Consequently, the contact areas of the
glass plates with the droplet and with the surfactant solution may
vary as the droplet elongates. All the three interfacial areas must be
taken into account to calculate the surface energy. The total surface
energy is
\begin{equation}
E_S= \sigma_{FI} A_{\cal C}+2\sigma_{FG}A_G+2\sigma_{IG}(A-A_G),
\end{equation}
where the three surface tensions between ferrofluid-immiscible fluid,
ferrofluid-glass, and surfactant solution-glass, are denoted by
$\sigma_{FI},\sigma_{FG}$, and $\sigma_{IG}$ respectively, $A_C$ and $A_G$
are defined below and the total area
of the sample is denoted by $A$. The factors of $2$ in the second and
third terms of the surface energy account for the two glass surfaces.
The area of the droplet-surfactant solution interface $A_{\cal C}$ is
given approximately by the circumference of $\cal C$ multiplied by the
arc length of the bulge
\begin{equation}
\label{AC}
A_C=2 \pi r_1 (1+{3\over 16}\epsilon^2)
\Delta {(\pi/2-\beta) \over {\cos \beta}}.
\end{equation}
We use $r_1$ here to calculate the circumference of the droplet because is
the radius observed during the experiment. To first order in the aspect ratio,
using $r_1$ or $r_2$ in equation~(\ref{AC}) yields the same result.
The droplet's contact area with the glass plates must be adjusted
to maintain a constant total volume of ferrofluid as the droplet elongates.
We approximate the volume of the bulging region by the circumference of
$\cal C$ multiplied by the projected area of the bulge. The contact area
$A_G$ must be adjusted so that $A_G \Delta$ changes by the negative of the
change in volume of the bulge. Thus we write
\begin{equation}
A_G=2\pi r_2^2 {\bigg [}1- {3 \over 32} {\bigg \{}{(\pi/2-\beta)\over
{\cos^2 \beta}}-{\tan \beta}{\bigg \}} {\Delta \over r_2}
\epsilon^2{\bigg ]}.
\end{equation}
Using $r_2$ instead of $r_1$ makes the above result exact for zero elongation.
The area of ferrofluid in contact with the glass plates decreases with
elongation for an acute contact angle because the volume of the fluid
contained in the outward bulge of the droplet increases and therefore the
fluid contained in the bulk of the droplet decreases. For obtuse contact
angles exactly the opposite happens for similar reasons.
To understand how the contact angle affects the magnetic energy,
consider the work done by the magnetic field as we change the contact
angle from $90^\circ$ to $\beta$ while keeping the volume of the
droplet constant. This work, divided by the circumference ,
must be independent of $r_1$ in the limit $r_1$ going to infinity,
since the magnetic field near the surface of the droplet will not depend
on $r_1$ in the large $r_1$ limit. The work done by the magnetic field
is the difference in energy between the straight edge droplet with
contact angle of $90^\circ$ and the bulging droplet with a contact
angle of $\beta$. The demagnetizing energy of the bulging droplet must
therefore have the same dependence on $\ln ({r_1/\Delta})$ as the
straight edge droplet or the difference in the demagnetizing
energies divided by the circumference will be proportional to $\ln
({r_1/\Delta})$ and will blow up in the large $r_1$ limit. Hence, the
demagnetizing energy for the bulging droplet must be identical to
equation~(\ref{expansion}) but with different values $\tilde{B}$ and
$\tilde{C}$ replacing the constants $B$ and $C$.
As the droplet bulges inward or outward the charges on the surface get
distributed over a larger area, decreasing the demagnetizing energy.
The constant $\tilde{B}$ therefore has a smaller value for a bulging
(inward or outward) droplet than $B$, the value for a straight-edged
droplet. However, since the demagnetizing energy is always positive,
smaller demagnetizing energy ($\tilde{B} < B$) implies a weaker
dependence of demagnetizing energy on elongation. Thus, we expect the
value of $\tilde{C}$ to be smaller for a bulging droplet than the
value $C$ for a straight-edged droplet.
Finally, consider how the contact angle dependence of surface and magnetic
energies affect the elongation calculated in equation~(\ref{elongation})
for the case $\beta=90^{\circ}$. The $\epsilon$ dependence of the
surface energy remains quadratic, but the coefficient now depends upon
a linear combination of the three surface tensions $\sigma_{FI}$,
$\sigma_{FG}$ and $\sigma_{IG}$. This combination will replace
$\sigma_{FI}$ in equation~(\ref{elongation}). The functional form of
the magnetic energy remains unchanged, but the values of $B$ and $C$
depend on contact angle. Thus the smaller value $\tilde{C}$ replaces $C$ in
equation~(\ref{elongation}). For $\beta \ne 90^\circ$ the experiment
cannot be used to determine $\sigma_{FS}$ unless $\sigma_{FG}$ and
$\sigma_{IG}$ are known. Since the $\beta$ in general is not
$90^\circ$, it is only possible to measure the effective surface
tension during elongation, and not $\sigma_{FI}$ itself.
\section{Conclusions}
We study the elongation ferrofluid droplets, confined in thin film geometry,
under weak applied field. Our theoretical calculations predict the elongation
of a droplet depends logarithmically on aspect ratio. This behavior
contrasts with the case of unconfined 3-dimensional droplets where elongation
is directly proportional to undeformed droplet radius. We measured the
elongation of ferrofluid droplets in an experiment performed on ferrofluid
droplets in a ferrofluid/water/surfactant emulsion. The results of our
experiment agree with the functional form our theoretical prediction,
however the experimentally measured value of $C$ differs from the
predicted value.We suggest the droplet contact angle with the confining
plates as a source of this discrepancy.
\acknowledgements
We acknowledge partial support for this research under NSF grant DMR-9732567.
|
\section{Introduction}
In recent years several molecular dynamics computer simulations have
been done in order to investigate the structure and dynamics of sodium
silicate melts and glasses (Smith, Greaves, and Gillan 1995, Cormack
and Cao 1997). By using the potential proposed by Vessal, Amini,
Fincham and Catlow (1989) these authors have found that the structure
of systems like, e.g., sodium disilicate (SDS) is characterized by a
microsegregation in which the sodium atoms form clusters of a few atoms
between bridged SiO$_4$ units. In order to see whether this somewhat
surprising result is reproduced also by a different model from the
one of Vessal {\it et al.}~(1989) we have performed simulations of SDS
using a different potential (discussed in detail below). In addition to
the investigation of the structure we also study the dynamical
properties of SDS in order to see whether the finite size effects that
have been observed in pure silica (Horbach, Kob, Binder, and Angell
1996, Horbach, Kob, and Binder 1999a, and Horbach, Kob, and Binder
1999b) are present in SDS as well.
\section{Model}
The model potential we use to describe the interactions between the
ions in SDS is the one proposed by Kramer, de Man, and van Santen
(1991) which is a generalization of the so called BKS potential (van
Beest, Kramer, and van Santen 1990) for pure silica. It has the
following functional form:
\begin{equation}
\phi(r)=
\frac{q_{\alpha} q_{\beta} e^2}{r} +
A_{\alpha \beta} {\rm e}^{-B_{\alpha \beta}r} -
\frac{C_{\alpha \beta}}{r^6}\quad \alpha, \beta \in
[{\rm Si}, {\rm Na}, {\rm O}].
\label{eq1}
\end{equation}
Here $r$ is the distance between an ion of type $\alpha$ and an ion of
type $\beta$. The values of the parameters $A_{\alpha \beta}, B_{\alpha
\beta}$ and $C_{\alpha \beta}$ can be found in the original
publication. The potential (\ref{eq1}) has been optimized by Kramer
{\it et al.}~for zeolites, i.e.~for systems that have Al ions in
addition to Si, Na and O. In that paper the authors used for silicon
and oxygen the {\it partial} charges $q_{{\rm Si}}=2.4$ and
$q_{{\rm O}}=-1.2$, respectively, whereas sodium was assigned its real
ion charge $q_{{\rm Na}}=1.0$. With this choice charge neutrality is
not fulfilled in systems like SDS. To overcome this problem we
introduced for the sodium ions a position dependent charge $q(r)$
instead of $q_{{\rm Na}}$,
\begin{equation}
q(r)= \left\{
\begin{array}{l@{\quad \quad}l}
0.6 \left( 1+
\ln \left[ C \left(r_{{\rm c}}-r \right)^2+1 \right] \right) &
r < r_{{\rm c}} \\
0.6 & r \geq r_{{\rm c}}
\end{array} \right. \label{eq2}
\end{equation}
which means that for $r \ge r_{{\rm c}}$ charge neutrality is valid
($q(r)=0.6$ for $r \ge r_{{\rm c}}$). Note that $q(r)$ is continuous at
$r_{{\rm c}}$. We have fixed the parameters $r_{{\rm c}}$ and $C$ such
that the experimental mass density of SDS and the static structure
factor from a neutron scattering experiment (see below) are reproduced
well. From this fitting we have obtained the values $r_{{\rm
c}}=4.9$\AA~and $C=0.0926$\AA$^{-2}$. With this choice the charge
$q(r)$ crosses smoothly over from $q(r)=1.0$ at $1.7$ \AA~to $q(r)=0.6$
for $r \ge r_{{\rm c}}$.
The simulations have been done at constant volume with the density of
the system fixed to $2.37 \, {\rm g}/{\rm cm}^3$. The equations of
motion were integrated with the velocity form of the Verlet algorithm
and the Coulombic contributions to the potential and the forces were
calculated via Ewald summation. The time step of the integration was
$1.6$ fs. In this paper we investigate the properties of SDS in the
liquid state at $T=2100$ K and in the glass state at $T=300$ K. The
equilibration time at $T=2100$ K was two million time steps thus
corresponding to a real time of $3.5$ ns. At this temperature we
simulated systems with $N=1008$ and $N=8064$ particles. In order to
improve the statistics two independent runs were done for the large
system and eight independent runs for the small system. The glass state
was produced by cooling the system from equilibrated configurations at
$T=1900$~K with a cooling rate of $1.16 \cdot 10^{12} \, {\rm K}/{\rm
s}$. The pressure is $4.5$ GPa at $T=2100$ K and $0.96$ GPa at $T=300$
K.
\section{Results}
In order to demonstrate that our model is able to reproduce the
structure of real SDS very well we compare the static structure factor
$S^{{\rm neu}}(q)$ with the one from a neutron scattering experiment by
Misawa, Price, and Suzuki (1980). To calculate $S^{{\rm neu}}(q)$ one
has to weight the partial structure factors from the simulation with the
experimental coherent neutron scattering lengths $b_{\alpha}$ ($\alpha
\in [{\rm Si, Na, O}]$):
\begin{equation}
S^{{\rm neu}}(q) =
\frac{1}{\sum_{\alpha} N_{\alpha} b_{\alpha}^2}
\sum_{kl} b_k b_l
\left< {\rm e}^{i {\bf q} \cdot [ {\bf r}_k - {\bf r}_l ]}
\right> . \label{eq3}
\end{equation}
The values for $b_{\alpha}$ are $0.4149 \cdot 10^{-12}$ cm, $0.363 \cdot
10^{-12}$ cm and $0.5803 \cdot 10^{-12}$ cm for silicon, sodium and
oxygen, respectively. They are taken from Susman, Volin, Montague, and
Price (1991) for silicon and oxygen and from Bacon (1972) for sodium.
Fig.~\ref{fig1} shows $S^{{\rm neu}}(q)$ from the simulation and the
experiment at $T=300$~K. We see that the overall agreement between
simulation and experiment is good. For $q > 2.3$ \AA$^{-1}$, which
corresponds to length scales of next nearest Si--O and Na--O neighbors,
the largest discrepancy is at the peak located at $q=2.8$~\AA$^{-1}$
where the simulation underestimates the experiment by approximately
$15$\% in amplitude. Very well reproduced is the peak at
$q=1.7$ \AA$^{-1}$, which is called the first sharp diffraction peak,
and which is a prominent feature in pure silica as well. In silica this
peak arises from the tetrahedral network structure since the
length scale which corresponds to it, i.e.~$2\pi/1.7 \; {\rm \AA}^{-1}=
3.7$ \AA, is approximately the spatial extent of two connected SiO$_4$
tetrahedra. From the figure we recognize that this structure is
partly present in SDS also. The peak at $q=0.95$ \AA$^{-1}$ is not
present in the experimental data which might be due to the fact that in
this $q$ region the experimental resolution is not sufficient. By
looking at the coordination number distributions, discussed below, we
see that the peak at $q=0.95$ \AA$^{-1}$ is related to
a super structure which is formed by the sodium and silicon atoms. In
agreement with this interpretation the length scale corresponding to
this peak, i.e.~$2 \pi/0.95 \; {\rm \AA}^{-1}=6.6$ \AA, is two times
the mean distance of nearest Na--Na or Na--Si neighbors.
The coordination number distribution $P_{\alpha \beta}(z)$ for
different pairs $\alpha \beta$ gives the probability that an ion of
type $\alpha$ has exactly $z$ nearest neighbors of type $\beta$. By
definition two neighboring atoms have a distance from each other which
is less than the location of the first minimum $r_{{\rm min}}$ of the
corresponding partial pair correlation function $g_{\alpha \beta}(r)$.
From the functions $g_{\alpha \beta}(r)$ we find for $r_{{\rm min}}$
the values $3.6$~\AA, $5.0$~\AA, $2.35$~\AA, $5.0$~\AA, $3.1$~\AA, and
$3.15$~\AA~for the Si--Si, Si--Na, Si--O, Na--Na, Na--O, and~O--O
correlations. We note that $P_{{\rm Si-O}}(z)$ is larger than $0.99$
for $z=4$ at $T=2100$ K and $T=300$ K which means that nearly every
silicon atom is four fold coordinated with oxygen atoms forming a
SiO$_4$ tetrahedron. Some of the distribution functions are shown in
Fig.~\ref{fig2}. We recognize from Fig.~\ref{fig2}a that about
$65$\% of the oxygen atoms are bridging oxygens between two tetrahedra
($P_{{\rm O-Si}}(z=2) \approx 0.65$), and that about $28$\% of the
oxygen atoms form dangling bonds ($z=1$) with corresponding silicon
atoms. In the neighborhood of these dangling bonds sodium atoms are
located. This means that the sodium atoms partly destroy the
(disordered) tetrahedral network which is the structure for pure
silica. A significant number of oxygen atoms have no silicon atoms, but
only sodium atoms as direct neighbors. In Fig.~\ref{fig2}b we show
$P_{{\rm Na-Na}}(z)$ at $T=2100$ K and $T=300$ K, and we see that
essentially the two distributions coincide with a
a mean value between $z=8$ and $z=9$. Basically the
same distribution is found for $P_{{\rm Na-Si}}(z)$. Therefore, every
sodium atom is surrounded by $8$--$9$ other sodium atoms and $8$--$9$
silicon atoms. Since the mean distance between Na--Na and Na--Si
neighbors is approximately the same, i.e.~about $3.3$ \AA, we can
conclude that sodium and silicon atoms form a spherical super structure
in which every sodium atom is surrounded by a first shell of
oxygen atoms and a second shell of silicon and sodium atoms. In
Fig.~\ref{fig2}b we have also included $P_{{\rm Na-O}}(z)$ and we
recognize that every sodium atom has on average about $4$--$5$ nearest
oxygen neighbors. There is again no essential difference between
$P_{{\rm Na-O}}$ for $T=2100$ K and $P_{{\rm Na-O}}$ for $T=300$ K,
although the pressure is about a factor $4.5$ higher at $T=2100$ K.
This means that the structural features which we observe at $T=2100$~K
are not formed due to the relatively high pressure. Nevertheless, we
emphasize that the small difference between $P(z)$ in the liquid and in
the glass state is partly due to the high cooling rate of about
$10^{12} \; {\rm K}/{\rm s}$ we used to produce the structures at
$T=300$ K. A careful analysis of the cooling rate effects for SDS will
be presented elsewhere (Horbach, Kob, and Binder 1999c).
Having described the structure of SDS we turn now our attention to
a dynamical quantity, namely the self part of the dynamic structure
factor $S_{{\rm s}} (q, \nu)$, which depends on frequency $\nu$ and the
magnitude of the wave--vector $q$. It is defined by
\begin{eqnarray}
S_{{\rm s}}(q,\nu) & = & \frac{N_{\alpha}}{N} \int_{-\infty}^{\infty}
dt \; {\rm e}^{-2 \pi \nu t}
\left<
{\rm e}^{i {\bf q} \cdot
[{\bf r}_{\alpha}(t) -
{\bf r}_{\alpha}(0)]} \right> \nonumber \\
& = & \frac{N_{\alpha}}{N} \int_{-\infty}^{\infty}
dt \; {\rm e}^{-2 \pi \nu t} F_{{\rm s}}(q,t)
\quad \quad \quad \quad
\alpha \in [{\rm Si, Na, O}]
\label{eq4}
\end{eqnarray}
where ${\bf r}_{\alpha}(t)$ is the position vector of a tagged particle
of type $\alpha$ at time $t$ and $F_{{\rm s}}(q,t)$ is the incoherent
intermediate scattering function. The details of the Fourier
transformation in (\ref{eq4}) are given elsewhere (Horbach {\it et
al.}~1999a). $S_{{\rm s}}(q,\nu)$ for silicon, sodium and oxygen is
shown in Fig.~\ref{fig3}a at the temperature $T=2100$ K and the
wave--vector $q=1.7$ \AA$^{-1}$ for the two system sizes $N=1008$ and
$N=8064$. For sodium the curves for the small and the large system
coincide over the whole frequency range. This is not the case for
silicon and oxygen for which $S_{{\rm s}}(q, \nu)$
has no system size dependence for frequencies $\nu \ge 1.1$~THz
whereas for smaller frequencies there is a missing of intensity for the
small system. The vibrational modes causing a shoulder around $\nu =
0.9$ THz in $S_{{\rm s}}(q, \nu)$ are usually called boson peak
excitations. In the small system a part of these excitations is missing
due to the loss of intensity for $\nu \le 1.1$ THz. An explanation of
this behavior for the case of silica, which shows qualitatively the
same finite size effects, can be found in Horbach {\it et al.} (1999a
and 1999b). Due to the sum rule $\int d \nu \; S_{{\rm s}}(q,\nu) = 1$
the missing of the boson peak excitations for frequencies 0.4 THz $\le
\nu \le$ 1.1 THz in the small system has to be ``reshuffled'' to
smaller frequencies leading to a broadening and an increase of the
quasielastic line around $\nu = 0$. Since the quasielastic line is
outside the frequency resolution of our Fourier transformation the
consequences in the change of the quasielastic line can be observed
better in the Fourier transform of $S_{{\rm s}}(q,\nu)$, i.e.~the
incoherent intermediate scattering function $F_{{\rm s}}(q,t)$, which
is shown in Fig.~\ref{fig3}b for the system sizes $N=8064$ and $N=1008$
at $q=1.7$ \AA$^{-1}$. We recognize from this figure that $F_{{\rm
s}}(q,t)$ shows a two step relaxation behavior similar to the case of
silica and fragile glassformers (Ngai, Riande, and Ingram 1998). As
expected from our results for $S_{{\rm s}}(q,\nu)$ the scattering
functions $F_{{\rm s}}(q,t)$ have no system size dependence for
sodium. In $F_{{\rm s}}(q,t)$ for silicon and oxygen the height of the
plateau increases and the $\alpha$ relaxation process shifts to longer
times with decreasing system size. Furthermore, the scattering
functions for the small system show a pronounced oscillation for
$t>0.2$ ps which is due to the fact that the boson peak excitations
present in the small system cause a peak in $S_{{\rm s}}(q,\nu)$
whereas in the large system only a shoulder is present. Finally we
mention that the finite size effects in the dynamics of SDS are found
in the whole $q$ range and, moreover, do not affect the static
properties of SDS.
{\bf Acknowledgments:}
This work was supported by SFB 262/D1 and by Deutsch--Israelisches
Projekt No.~352--101. We also thank the RUS in Stuttgart for a
generous grant of computer time on the Cray T3E.
\newpage
\section*{References}
\begin{trivlist}
\item[]
Bacon, G. E., 1972, {\it Acta Cryst. A}, {\bf 28}, 357.
\item[]
Cormack, A. N., and Cao, Y., in {\it Modelling of Minerals and
Silicated Materials}, Eds.: B. Silvi and P. Arco (Kluwer,
Dordrecht 1997).
\item[]
Horbach, J., Kob, W., Binder, K., and Angell C. A., 1996,
{\it Phys. Rev. E}, {\bf 54}, R5889.
\item[]
Horbach, J., Kob, W., and Binder, K., 1999a (submitted to Phys. Rev. B)
\item[]
Horbach, J., Kob, W., and Binder, K., 1999b, to appear in the
Proceedings on {\it Neutrons and Numerical Methods, Grenoble, 1998},
preprint in cond-mat/9901162
\item[]
Horbach, J., Kob, W., and Binder, K., 1999c (to be published)
\item[]
Kramer, G. J., de Man, A. J. M., and van Santen, R. A., 1991,
{\it J. Am. Chem. Soc.}, {\bf 113}, 6435.
\item[]
Misawa, M., Price, D. L., and Suzuki, K., 1980,
{\it J. Non--Cryst. Solids}, {\bf 37}, 85.
\item[]
Ngai, K. L., Riande, E., and Ingram, M.D., (editors), 1998,
{\it J. Non--Cryst. Solids}, {\bf 235--237} (Proceedings of the
{\it Third International Discussion Meeting on Relaxations in
Complex Systems}, Vigo, 1997).
\item[]
Smith, W., Greaves, G. N., and Gillan, M. J., 1995,
{\it J. Chem. Phys.}, {\bf 103}, 3091.
\item[]
Susman, S., Volin, K. J., Montague, D. G., and Price, D. L., 1991,
{\it Phys. Rev. B}, {\bf 43}, 11076.
\item[]
van Beest B. W., Kramer G. J., and van Santen R. A., 1990,
{\it Phys. Rev. Lett.}, {\bf 64} 1955.
\item[]
Vessal, B., Amini, A., Fincham, D., and Catlow, C. R. A., 1989,
{\it Philos. Mag. B}, {\bf 60}, 753.
\end{trivlist}
\newpage
\begin{figure}[f]
\psfig{file=fig1.eps,width=13cm,height=9.5cm}
\caption{Comparison of the static structure factor $S^{{\rm neu}}(q)$
from our simulation (solid line) with the experimental
data of Misawa {\it et al.}~(1980) (dashed line). \label{fig1}}
\end{figure}
\begin{figure}[f]
\psfig{file=fig2.eps,width=13cm,height=9.5cm}
\caption{Distributions of the coordination number for SDS at $T=300$ K
(open circles) and $T=2100$ K (filled circles),
a) $P_{{\rm O-Si}}(z)$, b) $P_{{\rm Na-O}}(z)$ and
$P_{{\rm Na-Na}}(z)$. \label{fig2}}
\end{figure}
\begin{figure}[f]
\psfig{file=fig3a.eps,width=13cm,height=9.5cm}
\psfig{file=fig3b.eps,width=13cm,height=9.5cm}
\caption{a) Self part of the dynamic structure factor for silicon,
sodium and oxygen for the system sizes $N=1008$ and
$N=8064$ at the temperature $T=2100$~K and the
wave--vector $q=1.7$ \AA$^{-1}$. The vertical line at
$\nu=1.1$ THz marks the frequency below which there is a loss
of intensity in the small system. b) Incoherent intermediate
scattering function $F_{{\rm s}}(q,t)$ under the same
conditions as in a).\label{fig3}}
\end{figure}
\end{document}
|
\section{Introduction and Results}
In the 90's, the celebrated KAM (Kolmogorov-Arnold-Moser) theory has been
successfully extended to infinite dimensional settings so as to deal with certain
classes of partial differential equations carrying a Hamiltonian structure,
including, as a typical example, wave equations of the form
\begin{equation}\label{1.1}
u_{tt}-u_{xx}+ V(x)u= f(u),\quad f(u)=O(u^2),\end{equation}
see Wayne \cite{Wa}, Kuksin \cite{K} and P\"oschel \cite{P1}. In such
papers, KAM theory for lower dimensional tori \cite{M1}, \cite{Me}, \cite{E}
(i.e., invariant tori of dimension lower than the number of degrees of freedom),
has been generalized in order to prove
the existence of small-amplitude quasi-periodic solutions for
(\ref{1.1}) subject to Dirichlet or Neumann boundary conditions (on a
finite interval for odd and analytic nonlinearities $f$). The technically
more difficult periodic boundary condition case has been later considered by
Craig and Wayne \cite{CW} who established the existence of periodic solutions.
The techniques used in \cite{CW} are based not on KAM theory, but rather
on a generalization of the Lyapunov-Schmidt procedure and on techniques by
Fr\"ohlich and Spencer \cite{FS}. Recently, Craig and Wayne's approach
has been significantly improved by Bourgain \cite{B1}, \cite{B2}
who obtained the existence of quasi-periodic solutions for certain kind
of 1D and, most notably, 2D partial differential equations with periodic boundary
conditions.
\smallskip\noindent
The technical reason why KAM theory has not been used to treat the periodic
boundary condition case is related to the multiplicity of the spectrum of the
Sturm-Liouville operator $A=-\frac{d^2}{dx^2}+V(x)$. Such multiplicity leads
to some extra ``small denominator"problems related to the so called normal
frequencies.
\smallskip\noindent
The purpose of this paper is to show that, allowing for more general normal
forms, one can indeed use KAM techniques to deal also with the multiple
normal frequency case arising in PDE's with periodic boundary conditions.
\smallskip\noindent
A rough description of our results is as follows.
Consider the periodic boundary problem for (\ref{1.1})
with an analytic nonlinearity $f$ and a real analytic (or smooth enough)
potential $V$. Such potential will be taken in a
$d$-dimensional family of functions parameterized by a real $d$--vector $\xi$,
$V(x)=V(x,\xi)$, satisfying
general non-degenerate (``non--\-resonance--\-of--\-eigenvalue")
conditions. Then for ``most" potentials in the family (i.e. for most $\xi$ in
Lebesgue measure sense), there exist small-amplitude
quasi-periodic solutions for (\ref{1.1}) corresponding to
$d$-dimensional KAM tori for the associated infinite dimensional
Hamiltonian system. Moreover (as usual in the KAM approach)
one obtains, for the constructed solutions, a
local normal form which provides linear
stability in the case the operator $A$ is positive definite.
\smallskip\noindent
We believe that the technique used in this paper can be generalized so
as to cover 2D wave equations.
\smallskip\noindent
The paper is organized as follows: In section 2 we formulate a general
infinite dimensional KAM Theorem designed to deal with multiple normal
frequency cases; in section 3 we show how to apply the preceding KAM Theorem
to the nonlinear wave equation (\ref{1.1}) with periodic boundary
conditions. The proof of the KAM Theorem is provided in sections 4$\div$6. Some
technical lemmata are proved in the Appendix.
\section{An infinite dimensional KAM Theorem}
In this section we will formulate a KAM Theorem in an infinite dimensional
setting which can be applied to some 1D partial differential equations with
periodic boundary conditions.
\smallskip\noindent
We start by introducing some notations.
\subsection{ Spaces}
For $n\in \Bbb N$, let $d_n\in \Bbb Z_+$ be positive {\sl even}
integers\footnote{We use the notations
$\Bbb N=\{0, 1,2,\cdots\}, \Bbb Z_+=\{1,2,\cdots\}$.}.
Let $\Cal Z\equiv \otimes_{n\in \Bbb N} {\Bbb C}^{d_n}$: the coordinates in
$\Cal Z$ are given by
$z=(z_0,z_1, z_2, \cdots)$ with
$z_n\equiv (z_n^1, \cdots, z_n^{d_n})\in {\Bbb C}^{d_n}$.
Given two real numbers $a, \rho$, we consider the
(Banach) subspace of $\Cal Z$ given by
\[\Cal Z_{a,\rho}=\{z\in \Cal Z: \ \ |z|_{a,\rho} <\infty \}\]
where the norm $|\cdot |_{a,\rho}$ is defined as
\[ |z|_{a,\rho}= |z_0|+\sum_{n\in {\Bbb Z}_+} |z_n|n^a e^{n\rho},\]
(and the norm in ${\Bbb C}^{d_n}$ is taken to be the 1--norm $|z_n|=
\sum_{j=1}^{d_n}|z_n^j|$).
\smallskip\noindent
{\sl In what follows, we shall consider either $a= 0$ and $\rho >0$ or $a>0$ and
$ \rho=0$
(corresponding respectively to the analytic case or the finitely smooth case)}.
\smallskip\noindent
The role of complex neighborhoods in phase space of KAM theory will be
played here by the set
\[ \Cal P_{a, \rho}\equiv \hat {\Bbb T}^d\times \Bbb C^d\times
\Cal Z_{a, \rho},\]
where $\hat{\Bbb T}^d$ is the complexification of the real torus
$\Bbb T^d= \Bbb R^d/2\pi\Bbb Z^d$.
\smallskip\noindent
For positive numbers $r, s$ we denote by
\begin{equation}
D_{a, \rho}(r, s)=\{(\theta,I,z)\in \Cal P_{a,\rho}:
|{\rm Im}\, \theta|<r, |I|<s^2,
|z|_{a,\rho}<s\}
\end{equation}
a complex neighborhood of $\Bbb T^d\times\{I=0\}\times\{ z=0\}$.
Finally, we denote by $\Cal O$ a given compact set in $\Bbb R^d$
with positive Lebesgue measure: $\xi\in \Cal O$ will parameterize a selected
family of potential $V=V(x, \xi)$ in (\ref{1.1}).
\subsection { Functions }
We consider functions $F$ on $D_{a,\rho}(r,s)\times \Cal O$ having the following
properties: (i) $F$ is real for real arguments; (ii) $F$ admits an expansion of the
form
\begin{equation} \label{Fa}
F=\sum_\alpha F_\alpha z^\alpha,\end{equation}
where the multi-index
$\alpha$ runs over the set
$\alpha\equiv (\alpha_0,\alpha_1,...)
\in \otimes_{n\in\Bbb N}\Bbb N^{d_n}$ with finitely many
non-vanishing components\footnote{Thus $\exists$ $n_0>0$ such that
$\displaystyle z^\alpha\equiv \prod_{n=0}^{n_0} z_n^{\alpha_n}\equiv
\prod_{n=0}^{n_0} \prod_{j=1}^{d_n} (z_n^j)^{\alpha_n^j}$.}
$\alpha_n$; (iii) for each $\alpha$, the
function
$F_\alpha=F_\alpha(\theta, I, \xi)$ is real analytic in the variables
$(\theta, I)\in
\{ |{\rm Im} \theta|<r, |I|<s^2\}$; (iv) for each $\alpha$, the dependence of
$F_\alpha$ upon the parameter $\xi$ is of class $C^{\bar d}_W(\Cal O)$ for
some $\bar d>0$ (to be fixed later): here $C^m_W(\Cal O)$ denotes the class of
functions which are
$m$ times differentiable on the closed set $\cal O$ in the sense of Whitney
\cite{Whit}.
\smallskip\noindent
The convergence of the expansion (\ref{Fa}) in $D_{a,\rho}(r,s)\times \Cal O$
will be
guaranteed by assuming the finiteness of the following weighted norm:
\begin{equation}\label{Fnorm}
\|F\|_{D_{a,\rho}(r,s),\Cal O}\equiv\sup_{|z|_{a,\rho}\le s}\sum_{\alpha}
\|F_\alpha\|\,\, |z^\alpha|
\end{equation}
where, if $\displaystyle F_\alpha=\sum_{k\in {\Bbb Z}^d, l\in {\Bbb N}^d} F_{kl\alpha}(\xi)I^l
e^{{\rm i}\langle k, \theta\rangle}\; $, $\|F_\alpha\|$ is defined as
\begin{equation} \label{Falpha}\|F_\alpha\|\equiv
\sum_{k, l}
|F_{kl\alpha}|_{\Cal O}\,
s^{2|l|} e^{|k|r},\qquad
|F_{kl\alpha}|_{\Cal O}\equiv
\max_{|p|\le \bar d^2}|\frac{\partial^pF_{kl\alpha}}
{\partial\xi^p}|,\end{equation}
(the derivatives with respect to $\xi$ are in the sense of Whitney).
\smallskip\noindent
The set of functions $F: D_{a,\rho}(r,s) \times \Cal O\to {\Bbb C}$ verifying
(i)$\div$(iv) above with finite
$\|\cdot\|_{ D_{a,\rho}(r,s),\cal O}$ norm
will be denoted by $\Cal F_{D_{a, \rho}(r,s), \Cal O}$.
\subsection {Hamiltonian vector fields and Hamiltonian equations}
To functions $F\in \Cal F_{D_{a, \rho}(r,s),\cal O}$, we associate a Hamiltonian
vector field defined as
\[X_F =(F_I, -F_\theta, \{{\rm i}J_{d_n}F_{z_n}\}_{n\in \Bbb N}),\]
where $J_{d_n}$ denotes the standard symplectic matrix
$
\left(\begin{array}{cc}
0 & I_{d_n/2}\\
-I_{d_n/2} & 0
\end{array} \right )$
and ${\rm i}=\sqrt{-1}$; the derivatives of $F$ are defined as the derivatives
term--by--term of the series (\ref{Fa}) defining $F$. The appearence of the imaginary
unit is due to notational convenience and will be justified later by the use of
complex canonical variables.
\smallskip\noindent
Correspondingly we consider the
Hamiltonian equations\footnote{Dot stands for
the time derivatives $d/dt$.}
\begin{equation}\label{2.5}
\dot \theta= F_I, \quad \dot I=-F_\theta,\quad
\dot z_n
= {\rm i}J_{d_n}F_{z_n},\quad n\in \Bbb N.
\end{equation}
{\sl A solution of such equation is intended to be just a $C^1$ map from an interval
to the domain of definition of $F$, $D_{a,\rho}(r,s)$, satisfying} (\ref{2.5}).
\smallskip\noindent
Given a real number $\bar a$, we define also a weighted norm for $X_ F$ by
letting\footnote{The norm $\|\cdot\|_{D_{a, \rho}( r,s), \cal O}$ for scalar
functions is defined in (\ref{Fnorm}). For vector (or matrix--valued) functions $G:
D_{a, \rho}( r,s)\times {\cal O}\to {\Bbb C}^m$, ($m<\infty$) is similarly defined as
$\|G\|_{D_{a, \rho}( r,s), \cal O}=\sum_{i=1}^m\|G_i\|_{D_{a, \rho}( r,s), \cal O}$
(for the matrix--valued case the sum will run over all entries).}
\begin{eqnarray}\label{weightednorm}
&&\|X_F\|_{\!{}_{D_{a, \rho}(r,s), \cal O}}^{\bar a,\rho}\equiv \\
\quad &&\|F_I\|_{\!{}_{D_{a, \rho}(r,s), \cal O}}+
\frac 1{s^2}\|F_\theta\|_{\!{}_{D_{a, \rho}(r,s), \cal O}}+
\frac 1s(\|F_{z_0}\|_{\!{}_{D_{a, \rho}(r,s), \cal O}}+
\sum_{n\in {\Bbb Z}_+} \|F_{z_n}\|_{\!{}_{D_{a, \rho}(r,s), \cal O}}n^{\bar a} e^{n\rho}).
\nonumber\end{eqnarray}
\noindent{\bf Notational Remark} In what follows,
only the indices $r, s$ and the set $\Cal O$ will change while $a, \bar a, \rho$
will be kept fixed, therefore we shall usually denote
$\|X_F\|_{D_{a,\rho}(r,s),\Cal O}^{\bar a,\rho}$ by $
\|X_F\|_{r,s,\Cal O}$, $D_{a, \rho}(r,s)$ by $D(r,s)$ and
$\Cal F_{D_{a, \rho}(r,s), \Cal O}$ by $\Cal F_{r,s,\Cal O}$.
\medskip\noindent
\subsection{Perturbed Hamiltonians and the KAM result}
The starting point will be a family of integrable Hamiltonians
of the form
\begin{equation}\label{hamN}
N =\langle \omega(\xi),I\rangle + \frac 12\sum_{n\in {\Bbb N}}\langle A_n(\xi)z_n, z_n\rangle ,
\end{equation}
where $\xi\in \Cal O$ is a parameter, $A_n$ is a $d_n\times
d_n$ real symmetric matrix and $\langle \cdot,\cdot\rangle $ is the standard inner product;
here the phase space ${\cal P}_{a,\rho}$ is endowed with the symplectic form
$\displaystyle dI\wedge d\theta + {\rm i} \sum_n \sum_{j=1}^{d_n/2} z_n^j \wedge d
z_n^{j+d_n/2}$.
\smallskip\noindent
{\sl For simplicity, we shall take, later, $\omega(\xi)\equiv \xi$}.
\smallskip\noindent
For each $\xi\in \Cal O$, the Hamiltonian equations of motion for $N$, i.e.,
\begin{equation}\label{unperturbed}
\frac {d\theta}{dt}= \omega,\quad \frac {dI}{dt}=0, \quad\frac {dz_n}{dt}
= {\rm i}J_{d_n}A_n z_n, \quad n\in \Bbb N,
\end{equation}
admit special solutions
$(\theta, 0, 0)\to (\theta+\omega t, 0, 0)$ corresponding to
an invariant torus in $\Cal P_{a, \rho}$.
\smallskip\noindent
Consider now the perturbed Hamiltonians
\begin{equation}\label{hamH}
H=N+P =\langle \omega(\xi),I\rangle + \frac 12
\sum_{n\in \Bbb N}\langle A_n(\xi)z_n, z_n\rangle + P(\theta,I,z, \xi)
\end{equation} with $P\in \Cal F_{r,s,\Cal O}$.
\smallskip\noindent
Our goal is to prove that, for most values of parameter
$\xi \in \Cal O$ (in Lebesgue measure
sense), the Hamiltonian $H=N+P$ still admits
an invariant torus provided $\|X_P\|$ is sufficiently
small.
\medskip\noindent
In order to obtain this kind of result {\sl we
make the following assumptions on $A_n$ and the perturbation $P$}.
\bigskip\noindent
\noindent
(A1) {\it Asymptotics of eigenvalues:}
There exist $\bar d\in \Bbb N, \delta>0$ and $b\ge 1$ such that
$d_n \le \bar d $ for all $n$, and
\begin{equation}\label{asymp1}
A_n= \lambda_n
\left(\begin{array}{cc}
0 & I_{d_n/2}\\
I_{d_n/2} & 0
\end{array}\right)
+ B_n,\quad B_n=O(n^{-\delta})
\end{equation}
where $\lambda_n$ are real and independent of $\xi$ while $B_n$ may
depend on $\xi$; furthermore, the behaviour of $\lambda_n$'s is assumed
to be as follows
\begin{equation} \label{asymp2}
\lambda_n= n^b+o(n^b), \quad\frac {\lambda_m-\lambda_n}{m^b-n^b}
= 1+ o(n^{-\delta}), \ \ \ n< m.\end{equation}
\bigskip\noindent
\noindent
(A2) {\it Gap condition:}\ There exists $\delta_1>0$ such that
\[{\rm dist}\left(\sigma(J_{d_i}A_{i}),
\sigma(J_{d_j}A_j)\right)> \delta_1>0,\quad \forall i\ne j\ ;\]
($\sigma(\cdot)$ denotes ``spectrum of $\cdot$").
\medskip\noindent
Note that, for large $i,j$, the {\it gap condition} follows
from the asymptotic property.
\bigskip\noindent
\noindent
(A3) {\it Smooth dependence on parameters:}
All entries of $B_n$ are $\bar d^2$ Whitney--smooth functions of $\xi$ with
$C^{\bar d^2}_W$-norm bounded by some positive constant $L$.
\bigskip\noindent
\noindent
(A4) {\it Non-resonance condition:}
\begin{equation}\label{ndc}
{\rm meas}\{\xi\in \Cal O:\quad
\langle k, \omega(\xi)\rangle (
\langle k, \omega(\xi)\rangle +\lambda(\xi))(\langle
k,\omega(\xi)\rangle +\lambda(\xi)+\mu(\xi))=0\}=0,
\end{equation}
for each $0\ne k\in \Bbb Z^d$
and for any $ \lambda, \mu\in \bigcup_{n\in \Bbb N}\sigma(J_{d_n}A_n)$; meas
$\equiv$ Lebesgue measure.
\bigskip\noindent
\noindent
(A5) {\it Regularity of the perturbation}:
The perturbation $P\in \Cal F_{D_{a, \rho}(r,s), \Cal O}$ is {\sl regular} in the
sense that $\|X_P\|_{\!{}_{D_{a, \rho}(r,s), \cal O}}^{\bar a,\rho}<\infty$
with $\bar a>a$. In fact, we assume that one of the following holds:
\[ (a)\;\;\; \rho>0,\quad \bar a>a=0; \quad (b)\;\;\; \rho=0,\quad \bar a>a>0,\]
(such conditions correspond, respectively, to analytic or smooth
solutions). In the case of $d=1$ (i.e., the periodic solution case) one can
allow $\bar a=a$.
\bigskip\noindent
Now we can state our KAM result.
\begin{Theorem}\label{KAM}
Assume that $N$ in (\ref{hamN}) satisfies (A1) - (A4) and $P$ is regular
in the sense of (A5) and let $\gamma>0$.
There exists a positive constant
$\epsilon=\epsilon(d, \bar d, b,\delta, \delta_1,\bar a-a, L, \gamma)$
such that if $\|X_P\|_{\!{}_{D_{a, \rho}(r,s), \cal O}}^{\bar
a,\rho}<\epsilon$, then the following holds true.
There exists a Cantor set
$\Cal O_\gamma\subset\Cal O$ with ${\rm meas}(\Cal O\setminus \Cal O_\gamma)
\to 0$ as $\gamma\to 0$, and two maps (real analytic in $\theta$ and Whitney
smooth in $\xi\in \Cal O$)
$$\Psi: T^d\times \Cal O_\gamma\to D_{a,\rho}(r,s)\subset \Cal P_{a,\rho},\ \ \ \
\tilde \omega:\Cal O_\gamma\to R^d,$$
such that for any $\xi\in \Cal O_\gamma$ and $\theta\in T^d$
the curve $t\to \Psi(\theta+\tilde \omega(\xi) t,\xi)$ is a quasi-periodic
solution of the Hamiltonian equations governed by $H=N+P$. Furthermore,
$\Psi({\Bbb T}^d, \xi)$ is a smoothly embedded $d$-dimensional $H$-invariant
torus in $\Cal P_{a,\rho}$.
\end{Theorem}
\medskip\noindent
\noindent
{\bf Remarks}
(i) For simplicity we shall in fact assume that {\sl all eigenvalues $\lambda_i$ of
$A_n$ are positive for all $n$'s}. The case of some non positive eigenvalues can be
easily dealt with at the expense of a (even) heavier notation.
\medskip\noindent
(ii)
In the above case (i.e. positive eigenvalues), Theorem 1 yields {\sl linearly
stable} KAM tori.
\medskip\noindent
\noindent
(iii) The parameter $\gamma$ plays the role of the Diophantine constant for
the frequency $\tilde \omega$ in the sense that,
there is $\tau>0$ such that $\forall k\in {\Bbb Z}^d\backslash \{0\},$
\[\langle k, \tilde \omega\rangle
>\frac {\gamma}{2|k|^\tau}.\]
Notice also that $\Cal O_\gamma$ is claimed to be nonempty and big only for
$\gamma$ small enough.
\medskip\noindent
\noindent
(iv)
The regularity property $\bar a>a$ is needed for the measure estimates on $\Cal
O\backslash \Cal O_\gamma$.
As we already mentioned, such regularity requirement
is not necessary for periodic solution case, i.e., $d=1$. Thus
{\sl the above theorem applies immediately to the construction of periodic
solutions for nonlinear Schr\"odinger equations}.
\medskip\noindent
\noindent
(v). The non-degeneracy condition (\ref{ndc}) (which
is stronger than Bourgain's non-degenerate condition \cite{B2} but weaker than
Melnikov's one \cite{Me}) covers the multiple normal frequency case: this is the
technical reason that allows to treat PDE's with periodic boundary conditions.
\section{Application to 1D wave equations}
\noindent
In this section we show how Theorem \ref{KAM} implies the existence
of quasi-periodic solutions for 1D wave
equations with periodic boundary conditions.
\smallskip\noindent
Let us rewrite the wave equation (\ref{1.1}) as follows
\begin{eqnarray}\label{nonlinearwave}
& & u_{tt}+Au= f(u),\ \ \
Au\equiv
-u_{xx}+V(x,\xi)u, \ x, t\in {\Bbb R}, \nonumber\\
& & u(t,x)=u(t, x+2\pi),\quad u_t(t,x)=u_t(t, x+2\pi),
\end{eqnarray}
where $V(\cdot,\xi)$ is a {\it real--analytic}
({\it or smooth}) periodic potential
parameterized by some $\xi\in {\Bbb R}^d$ (see below) and
$f(u)$ is a {\it real--analytic}
function near $u=0$ with $f(0)=f'(0)=0$.
\smallskip\noindent
As it is well known, the operator $A$
with periodic boundary conditions admits an orthonormal basis of
eigenfunctions $\phi_n\in L^2({\Bbb T}), n\in {\Bbb N}$, with corresponding eigenvalues
$\mu_n$ satisfying the following asymptotics for large $n$
\[\mu_{2n-1}, \mu_{2n}= n^2+\frac 1{2\pi}
\int_{\Bbb T} V(x)dx+O(n^{-2}).\]
\smallskip\noindent
{\sl For simplicity, we shall consider the case
of vanishing mean value of the potential $V$ and assume that all eigenvalues are
positive:}
\begin{equation}\label{assum}
\int_{\Bbb T} V(x)dx=0\ ,\qquad
\mu_n\equiv \lambda_n^2>0\ ,\quad\forall\ n.
\end{equation}
\smallskip\noindent
Following Kuksin \cite{K} and Bourgain \cite{B1}, we consider a family of
real analytic (or smooth) potentials
$V(x, \xi),$ where {\sl the $d$-parameters
$\xi=(\xi_1, \cdots, \xi_d)\in \Cal O\subset {\Bbb R}^d$ are simply taken
to be a given set of $d$ frequencies $\lambda_{n_i}\equiv\sqrt{\mu_{n_i}}$}:
\begin{equation}\label{xml}
\xi_i\equiv\sqrt{\mu_{n_i}}\equiv\lambda_{n_i}\ , \qquad i=1,
\cdots, d
\end{equation}
where $\mu_{n_i}$ are (positive) eigenvalues of\footnote{Plenty of such potentials
may be constructed with, e.g., the inverse spectral theory.}
$A$.
\smallskip\noindent
We may also (and shall) require that there exists a positive $\delta_1>0$ such that
\begin{equation}\label{del1}
|\mu_k-\mu_h|> \delta_1\ ,
\end{equation}
for all $k>h$ except when $k$ is even and $h=k-1$ (in which case $\mu_k$ and $\mu_h$
might even coincide).
\smallskip\noindent
Notice that, in particular, having $d$ eigenvalues as {\it independent}
parameters excludes the constant potential case $V\equiv$ constant (where, of course,
all eigenvalues are double: $\mu_{2j-1}=\mu_{2j}=j^2+V$).
In fact, this case seems difficult to be handled by KAM
approach even in the finite dimensional case. Such difficulty does not arise,
instead, in the remarkable alternative approach developed by Craig, Wayne
\cite{CW} and Bourgain \cite{B1}, \cite{B2}.
\smallskip\noindent
Equation (\ref{nonlinearwave}) may be rewritten as
\begin{equation}\label{HH}
\dot u=v, \qquad \dot v +Au= f(u),
\end{equation}
which, as well known, may be viewed as the (infinite dimensional) Hamiltonian
equations $\dot u = H_v$, $\dot v=- H_u$ associated to the Hamiltonian
\begin{equation}
H=\frac 12( v,v)+\frac 12( Au,u)+\int_{\Bbb T} g(u)\ dx,
\end{equation}
where $g$ is a primitive of $(-f)$
(with respect to the $u$ variable) and $( \cdot,\cdot)$ denotes the scalar product
in $L^2$.
\smallskip\noindent
As in \cite{P1}, we introduce coordinates $q=(q_0, q_1, \cdots),
p=(p_0, p_1, \cdots)$ through the relations
\[u(x)=\sum_{n\in {\Bbb N}}\frac{q_n}{\sqrt{\lambda_n}}\phi_n(x),
\quad v=\sum_{n\in{\Bbb N}} \sqrt{\lambda_n}p_n\phi_n(x),
\]
where\footnote{Recall that, for simplicity, we assume that all eigenvalues $\mu_n$
are positive.}
$\lambda_n\equiv \sqrt{\mu_n}$. System (\ref{HH}) is then formally equivalent to
the lattice Hamiltonian equations
\begin{equation}\label{hameq} \dot q_n=\lambda_n p_n,
\quad \dot p_n=-\lambda_nq_n- \frac{\partial G}{\partial q_n},
\quad G\equiv \int_{\Bbb T}
g(\sum_{n\in {\Bbb N}}\frac{q_n}{\sqrt{\lambda_n}}\phi_n)dx\ ,\end{equation}
corresponding to the Hamiltonian function
$H= \sum_{n\in {\Bbb N}}\lambda_n(q_n^2+p_n^2)+ G(q)$.
Rather than discussing the above formal equivalence, we shall, following
\cite{P1}, use the following elementary observation (proved in the Appendix):
\begin{Proposition}\label{prop}
Let $V$ be analytic (respectively, smooth), let $I$ be an interval and let
$$
t\in I\to (q(t), p(t))\equiv \Big(\{q_n(t)\}_{n\ge 0},\{p_n(t)\}_{n\ge 0}\Big)
$$
be an analytic (respectively, smooth\footnote{Regularity refers to the components
$q_n$ and $p_n$.}) solution of (\ref{hameq}) such
that
\begin{equation}\label{reg}
\sup_{t\in I} \sum_{n\in \Bbb N} \Big( |q_n(t)|+|p_n(t)|\Big) \ n^a\ e^{n
\rho}<\infty
\end{equation}
for some $\rho>0$ and $a=0$ (respectively, for $\rho=0$ and $a$ big enough).
Then
\[u(t, x)\equiv\sum_{n\in {\Bbb N}}\frac{q_n(t)}{\sqrt{\lambda_n}}\phi_n(x),\]
is an analytic (respectively, smooth) solution of (\ref{nonlinearwave}).
\end{Proposition}
\smallskip\noindent
Before invoking Theorem 1 we still need some manipulations.
We first switch to complex variables:
$w_n=\frac{1}{\sqrt{2}}(q_n+{\rm i}p_n), \bar
w_n=\frac{1}{\sqrt{2}}(q_n-{\rm i}p_n)$.
Equations (\ref{hameq}) read then
\begin{equation}\label{hamw} \dot w_n=-{\rm i}\lambda_n w_n- {\rm i}
\frac{\partial \tilde G}{\partial \bar w_n}, \quad
\dot{\bar w}_n={\rm i}\lambda_n\bar w_n+ {\rm i}
\frac{\partial \tilde G}{\partial w_n},
\end{equation}
where the perturbation ${\tilde G}$ is given by
\begin{equation} \label{tildeG}\tilde G(w)=\int_{\Bbb T}
g(\sum_{n\in {\Bbb N}}\frac{w_n+\bar w_n}{\sqrt{2 \lambda_n}}\phi_n)dx.
\end{equation}
\smallskip\noindent
Next we
introduce standard action-angle variables $(\theta, I)=((\theta_1,\cdots,
\theta_d),( I_1, \cdots, I_d))$ in the
$(w_{n_1}, \cdots, w_{n_d},\bar w_{n_1}, \cdots, \bar w_{n_d} )$-space by
letting,
\[I_i= w_{n_i}\bar w_{n_i}, \quad i=1, \cdots, d,\]
so that the system (\ref{hamw}) becomes
\begin{eqnarray}\label{sysP}
\frac{d\theta_j}{dt}&=&\omega_j + P_{I_j},\quad
\frac {dI_j}{dt}=- P_{\theta_j},\quad j=1, \cdots, d,
\nonumber\\
\frac {dw_n}{dt}&=&-{\rm i}\lambda_nw_n-{\rm i} P_{\bar w_n},\ \ \
\frac {d\bar w_n}{dt}={\rm i}\lambda_n\bar w_n +{\rm i} P_{w_n},
\ n\ne n_1, n_2,\cdots, n_d,
\end{eqnarray}
where $P$ is just $\tilde G$ with the
$(w_{n_1}, \cdots, w_{n_d},\bar w_{n_1}, \cdots, \bar w_{n_d} )$-variables
expressed in terms of the $(\theta,I)$ variables and the {\sl frequencies}
$\omega=(\omega_1,...,\omega_d)$ coincide with the parameter $\xi$ introduced in
(\ref{xml}):
\begin{equation}\label{oml}
\omega_i\equiv \xi_i=\lambda_{n_i}\ .
\end{equation}
The Hamiltonian associated to (\ref{sysP}) (with respect to the
symplectic form
$dI\wedge d\theta+{\rm i}\sum_n dw_n \wedge d\bar w_n$) is given by
\begin{equation}\label{hamI}
H=\langle \omega, I\rangle +\sum_{n\ne n_1, \cdots, n_d}\lambda_nw_n\bar w_n+
P(\theta, I, w, \bar w, \xi),
\end{equation}
\medskip\noindent\noindent
{\bf Remark} Actually, in place of $H$ in (\ref{hamI}) one should consider the
{\sl linearization} of $H$ around a given point $I_0$ and let $I$ vary in a
small ball $B$ (of radius $0<s\ll |I_0|$) in the ``positive'' quadrant
$\{I_j>0\}$. In such a way the dependence of $H$ upon $I$ is obviously
analytic. For notational convenience we shall however do not report explicitly
the dependence of $H$ on $I_0$.
\medskip\noindent\noindent
Finally, to put the Hamiltonian in the form (\ref{hamH})
we couple the variables $(w_n, \bar w_n)$ corresponding
to ``closer" eigenvalues. More precisely, we let
$z_n=(w_{2n-1}, w_{2n}, \bar w_{2n-1},
\bar w_{2n})$ for large\footnote{Compare (A1).} $n$, say $n> \bar n> n_d$ and
denote by
$\displaystyle
z_0=(\{w_n\}_{ {0\le n\le \bar n}\atop{n\neq n_1,...,n_d}},
\{\bar w_n\}_{ {0\le n\le \bar n}\atop{n\neq n_1,...,n_d}})$
the remaining conjugated variables. The Hamiltonian (\ref{hamI}) takes the form
\begin{equation}
\label{hamiltonian z}H
=\langle \omega, I\rangle + \frac 12\sum_{n\in {\Bbb N}} \langle A_n z_n, z_n\rangle +
P(\theta, I, z, \xi),
\end{equation}
where
\begin{eqnarray}
A_n&=&{\rm Diag}(\lambda_{2n-1},\lambda_{2n},\lambda_{2n-1},\lambda_{2n}) \pmatrix{ 0
& I_2 \cr I_2 & 0 \cr}\nonumber\\
&=& \lambda_{2n} \pmatrix{ 0 & I_2 \cr I_2 & 0 \cr} +
\pmatrix{0&0&\lambda_{2n-1}-\lambda_{2n}&0\cr
0&0&0&0\cr
\lambda_{2n-1}-\lambda_{2n}&0&0&0\cr0&0&0&0\cr}
\ ,\nonumber
\end{eqnarray}
for $n>n_d$, while $A_0={\rm Diag}(\{\lambda_n\},$
$ \{\lambda_n\}; 1\le n\le n_d, n\ne n_1, \cdots,
n_d) \pmatrix{ 0 & I_{d_0} \cr I_{d_0} & 0 \cr}$ with $d_0=\bar n +1 -d$.
\smallskip\noindent
The perturbation $P$ in (\ref{hamiltonian z}) has the following (nice) regularity
property.
\begin{Lemma}\label{regularity}
Suppose that $V$ is real analytic in $x$ (respectively, belongs to the
Sobolev space $H^k({\Bbb T})$ for some $k\in {\Bbb N}$).
Then for small enough $\rho>0$ (respectively, $a>0$), $r>0$ and $s>0$
one has
\begin{equation} \|X_P\|_{\!{}_{D_{a, \rho}(r,s), \cal O}}^{a+1/2,\rho}=
O(|z|_{a,\rho}^2) \ ;
\label{Pz}
\end{equation}
here the parameter $a$ is taken to be 0 (respectively, the parameter $\rho$ is taken
to be 0).
\end{Lemma}
A proof of this lemma is given in the Appendix. In fact, $X_P$
is even more ``regular" (a fact, however, not needed in what follows):
(\ref{Pz}) holds with 1 in place of 1/2.
\medskip\noindent
The Hamiltonian (\ref{hamiltonian z}) is seen to satisfy all the assumptions of
Theorem \ref{KAM} with:
$d_n=4,n\ge 1;$ $d_0=\bar n +1 -d$; $\bar d=\max\{d_0,
4\}$; $b=1$; $\delta=2$;
$\delta_1$ choosen as in (\ref{del1});
$\bar a-a=\frac 12$.
Thus Theorem \ref{KAM} yields the following
\begin{Theorem}
Consider a family of 1D nonlinear wave equation (\ref{nonlinearwave})
parameterized by $\xi\equiv \omega\in \Cal O$ as above with $V(\cdot, \xi)$
real--analytic (respectively, smooth).
Then for any $0<\gamma\ll 1$, there is a subset $\Cal O_\gamma$ of $\Cal O$ with
${\rm meas}(\Cal O\backslash \Cal O_\gamma)\to 0$ as $\gamma\to 0$,
such that
${\rm (\ref{nonlinearwave})}_{\xi\in \Cal
O_\gamma}$ has a family of small-amplitude (proportional to
some power of $\gamma$), analytic (respectively, smooth)
quasi-periodic solutions of the form
\[ u(t, x)=\sum_n u_n(\omega_{1}'t, \cdots, \omega_{d}' t)
\phi_n(x)\]
where $u_n: {\Bbb T}^d\to {\Bbb R} $ and $\omega_{1}',\cdots, \omega_{d}'$
are close to $\omega_{1},\cdots, \omega_{d}$.
\end{Theorem}
\noindent
{\bf Remark}
As mentioned above, our KAM theorem (which applies only
to the case that not all the eigenvalues are multiple\footnote{
Recall that we require that the torus frequencies are independent parameters.}
and under the hypothesis that all $\mu_n$'s are
positive) implies that the quasi-periodic solutions obtained are {\it linearly
stable}. In the case that all the eigenvalues are double (as in the constant
potential case), one should not expect linear stability (see the example given by
Craig, Kuksin and Wayne \cite{CKW}).
We also notice that, essentially with only notational changes, the proof of the above
theorem goes through in the case that some of the eigenvalues are negative.
\section{ KAM step}
Theorem 1 will be proved by a KAM iteration which involves an infinite
sequence of change of variables.
\smallskip\noindent
At each step of the KAM scheme, we consider a Hamiltonian vector
field with
$$ H_\nu=N_\nu+P_\nu,$$
where $N_\nu$ is an ``integrable normal form" and
$P_{\nu}$ is
defined in some set of the form\footnote{Recall the notations from Section 2.}
$D(s_\nu, r_\nu)\times \Cal O_{\nu} $.
\smallskip\noindent
We then construct a map\footnote{Recall that the parameters $a$, $\rho$ and $\bar a$
are fixed throughout the proof and are therefore omitted in the notations.}
$$\Phi_\nu: D (s_{\nu+1}, r_{\nu+1})\times\Cal O_{\nu+1}
\subset D(r_\nu, s_\nu)\times\Cal O_\nu \to D(r_\nu, s_\nu)
\times\Cal O_{\nu}
$$
so that the vector field $X_{H_\nu
\circ\Phi_\nu}$ defined on $D(r_{\nu+1}, s_{\nu+1})$
satisfies
\[\|X_{H_\nu\circ\Phi_\nu}-X_{N_{\nu+1}}\|_{r_{\nu+1}, s_{\nu+1},
\Cal O_{\nu+1}}\le \epsilon_{\nu}^\kappa\]
with some new normal form $N_{\nu +1}$ and for some fixed $\nu$-independent
constant $\kappa>1$.
\smallskip\noindent
To simplify notations, in what follows, the quantities without subscripts refer to
quantities at the $\nu^{\rm th}$ step, while the quantities with subscripts $+$
denotes the corresponding quantities at the $(\nu+1)^{\rm th}$ step.
Let us then consider the Hamiltonian
\begin{equation} H=N+P\equiv e+ \langle \omega,I\rangle
+\frac 12\sum_{n\in {\Bbb N}} \langle A_nz_n, z_n\rangle
+P,
\label{3.1}
\end{equation}
defined in $D(r, s)\times\Cal O$; teh $A_n$'s are symmetric matrices. We
assume that
$\xi\in \Cal O$ satisfies\footnote
{
The tensor product (or direct product) of two $m\times n$, $k\times l$
matrices $A=(a_{ij}), B$
is a $(mk)\times (nl)$ matrix
defined by
\[ A\otimes B=(a_{ij}B)= \left(\begin{array}{ccc}
a_{11}B & \cdots & a_{1n}B\\
\cdots & \cdots & \cdots \\
a_{n1}B & \cdots & a_{mn}B
\end{array}\right)
.\]
$\|\cdot \|$ for matrix denotes the operator norm, i.e.,
$\|M\|=\sup_{|y|=1}|My|.$
Recall that $\omega$ and the $A_i$'s depend on $\xi$.
}
(for a suitable $\tau>0$ to be specified later)
\begin{eqnarray}\label{ODC}
& &|\langle k,\omega\rangle ^{-1} |< \frac{|k|^\tau}{\gamma} ,\;\;\;\;
\|(\langle k,\omega\rangle I_{d_i}+
{ A_iJ_{d_i}})^{-1}\|<(\frac{|k|^\tau}{\gamma})^{\bar d}
\nonumber\\
& &\|(\langle k, \omega\rangle I_{d_id_j}+(A_iJ_{d_i})
\otimes I_{d_j}-I_{d_i}\otimes (J_{d_j}A_j))^{-1}\|
<(\frac{|k|^\tau}{\gamma})^{\bar d^2},
\end{eqnarray}
\smallskip\noindent
We also assume that
\begin{equation}\label{aijM}
\max_{|p|\le \bar d^2}\|\frac{\partial^p A_n}{\partial\xi^p}\|\le
L,
\end{equation}
on $\Cal O$, and
\begin{equation}\|X_P\|_{ r,s,\Cal O}
\le \epsilon.
\label{3.3}
\end{equation}
We now let $0<r_+<r$, and define
\begin{equation}\label{+}
s_+=\frac 12s\epsilon^{\frac 13}, \quad
\epsilon_+=\gamma^{-{\rm c}}\Gamma(r-r_+)
\epsilon^{\frac 43},
\label{3.4-}
\end{equation}
where
$$\Gamma(t)\equiv \sup_{u\ge 1} u^{\rm c} e^{-\frac 14ut}
\sim t^{-{\rm c}}.$$
for $t>0$. Here and later, the letter $c$ denotes suitable (possibly different)
constants that do not depend on the iteration step\footnote{Actually, here
$c=\bar d^4\tau+\bar d^2\tau+\bar d^2+1$.}.
\smallskip\noindent
We now describe how to construct a set $\Cal O_+\subset \Cal O$ and
a change of variables
$\Phi: D_+\times\Cal O_+=D(r_+, s_+)\times \Cal O_+\to D(r,s)\times \Cal O$,
such that the transformed Hamiltonian
$H_+=N_++P_+\equiv H\circ \Phi$ satisfies
all the above iterative assumptions with new parameters
$s_+, \epsilon_+, r_+, \gamma_+, L_+$ and with $\xi\in \Cal O_+$.
\medskip\noindent
\subsection{Solving the linearized equation}
\smallskip\noindent
Expand $P$ into the Fourier-Taylor series
$$P=\sum_{k,l,\alpha} P_{kl\alpha}e^{{\rm i}\langle k, \theta\rangle}\; I^lz^\alpha$$
where $k\in {\Bbb Z}^d, l\in {\Bbb N}^d$ and $\alpha\in \otimes_{n\in {\Bbb N}}\Bbb N^{d_n}$
with finite many non-vanishing components.
\smallskip\noindent
Let $R$ be the truncation of $P$ given by
\begin{eqnarray}
R(\theta,I,z)&\equiv &P_0+P_1+P_2
\equiv \sum_{k,|l|\le 1} P_{kl0}e^{{\rm i}\langle k, \theta\rangle}\; I^l\nonumber\\
& &+\sum_{k,|\alpha|=1} P_{k0\alpha}e^{{\rm i}\langle k, \theta\rangle}\; z^\alpha+\sum_{
k,|\alpha|=2}
P_{k0\alpha} e^{{\rm i}\langle k, \theta\rangle}\; z^\alpha,\label{Pi}\end{eqnarray}
with
\[2|l|+|\alpha|=2\sum_{j=1, \cdots, d}
l_j+
\sum_{j\in {\Bbb N}} |\alpha_j| \le 2.\]
\smallskip\noindent
It is convenient to rewrite $R$ as follows
\begin{eqnarray}\label{R}
R(\theta, I, z)&=& \sum_{k,|l|\le 1} P_{kl0}e^{{\rm i}\langle k, \theta\rangle}\; I^l\nonumber\\
&+&\sum_{k,i} \langle R^{k}_i,z_i\rangle e^{{\rm i}\langle k, \theta\rangle}\; +\sum_{k,i,j}
\langle { R}^k_{ji}z_i, z_j\rangle e^{{\rm i}\langle k, \theta\rangle}\; ,
\end{eqnarray}
where $R^k_i,R^k_{ji}$ are respectively the $d_i$ vector and $(d_j\times d_i)$
matrix defined by
\begin{equation}\label{R_k}
R^{k}_i=\int\frac{\partial P}{\partial z_i}e^{-{\rm i}\langle k,\theta
\rangle}d\theta |_{z=0,I=0}, \ \ \
R^k_{ji}=
\frac {1+\delta_i^j}2\int\frac{\partial^2 P}{\partial z_j\partial z_i}
e^{-{\rm i}\langle k,\theta
\rangle}d\theta|_{z=0, I=0}.
\end{equation}
Note that $R^k_{ij}=(R^k_{ji})^T$.
\smallskip\noindent
Rewrite $H$ as
$ H=N+R+(P-R)$. By the choice of $s_+$ in (\ref{+}) and by the definition of
the norms, it follows immediately that
\begin{equation}
\label{P-R}
\|X_R\|_{r,s,\Cal O}\le
\|X_P\|_{r,s,\Cal O}\le\epsilon.
\end{equation}
Moreover $s_+, \epsilon_+$ are such that, in a smaller domain $D(r, s_+)$, we have
\begin{equation}
\label{P-R+}
\|X_{P-R}\|_ {r, s_+} \;<\;{\rm c}\; \epsilon_+.
\end{equation}
\smallskip\noindent
Then we look for a special $F$,
defined in domain $D_+=D(r_+, s_+),$
such that the time one map $\phi^1_F$ of the Hamiltonian vector field $X_F$
defines a map from $D_+\to D$ and transforms
$H$ into $H_+$.
\smallskip\noindent
More precisely, by second order Taylor formula, we have
\begin{eqnarray}
H\circ \phi^1_F &=&(N+R)\circ \phi_F^1+(P-R)\circ \phi^1_F\nonumber\\
&=& N+ \{N,F\}+R\nonumber\\
& &\ \ \ \ +\frac 12\int_0^1ds\int_0^s\{\{N+R,F\},F\}\circ \phi_F^{t}dt+\{R,F\}
+(P-R)\circ \phi^1_F.
\nonumber\\
&=& N_++P_+\nonumber\\
& & + \{N,F\}+R-P_{000}-\langle \omega', I\rangle -\sum_{n\in {\Bbb N}}
\langle R^0_{nn}z_n, z_n\rangle ,
\label{2.3}
\end{eqnarray}
where
\[
\omega'= \int\frac{\partial P}{\partial I}d\theta|_{I=0, z=0},\quad
R^0_{nn}=\int \frac{\partial^2 P}{\partial z_n^2}d\theta|_{I=0, z=0},
\]
$$N_+= N+P_{000}+\langle \omega', I\rangle +
\sum_{n\in {\Bbb N}}\langle R^0_{nn}z_n, z_n\rangle
$$
$$P_+=\frac 12\int_0^1ds\int_0^s\{\{N+R,F\},F\}\circ X_F^{t}dt+\{R,F\}
+(P-R)\circ \phi^1_F.$$
\medskip\noindent
We shall find a function $F$ of the form
\begin{eqnarray}\label{F}
F(\theta, I,z)&=& F_0+ F_1+F_2=\sum_{|l|\le 1,|k|\ne 0}
F_{kl0}e^{{\rm i}\langle k, \theta\rangle}\;
I^l+
\sum_{i\in {\Bbb N}} \langle F^k_i,z_i\rangle e^{{\rm i}\langle k, \theta\rangle}\; \nonumber\\
& &+\sum_{|k|+|i-j|\ne 0} \langle { F}^k_{ji}z_i, z_j\rangle e^{{\rm i}\langle k, \theta\rangle}\; ,
\label{3.14}
\end{eqnarray}
satisfying the equation
\begin{equation}\label{hom}
\{N,F\}+R-
P_{000}-\langle \omega', I\rangle -\sum_{n\in {\Bbb N}}\langle R^0_{nn}z_n, z_n\rangle =0.
\end{equation}
\begin{Lemma}\label{Fdef} Equation (\ref{hom}) is equivalent to
\begin{eqnarray}\label{Feq}
F_{kl0}&=& ({\rm i}\langle k,\omega\rangle )^{-1} P_{kl0}, \quad \ k\ne 0, |l|\le 1,
\nonumber\\
(\langle k,\omega\rangle I_{d_i} &+& { A_{d_i}J_{d_i}})F^k_i={\rm i} R^k_i,\nonumber\\
(\langle k,\omega\rangle I_{d_i}&+& A_{d_i}J_{d_i})F^k_{ij}- F^k_{ij}
( J_{d_j}A_j)={\rm i} R^k_{ij},\ \, |k|+|i-j|
\ne 0.
\end{eqnarray}
\end{Lemma}
\Proof Inserting $F$, defined in (\ref{F}), into (\ref{hom}) one sees that
(\ref{hom}) is equivalent to
the following equations\footnote{Recall the definition of $P_i$ in
(\ref{Pi}).}
\begin{eqnarray}
\label{Fcoefficents}
& &\{N, F_0\}+ P_0-\langle \omega', I\rangle =0, \nonumber\\
& & \{N, F_{1}\}+ P_{1}=0,\nonumber\\
& &\{N, F_{2}\}+ { P}_{2}-\sum_{n\in Z}\langle R^0_{nn}z_n, z_n\rangle =0.
\end{eqnarray}
The first equation in (\ref{Fcoefficents}) is obviously equivalent, by
comparing the coefficients, to the first equation in (\ref{Feq}).
To solve $\{N, F_1\}+ P_1=0$, we note that\footnote{Recall the definition
of $N$ in (\ref{3.1}).}
\begin{eqnarray}
\{N, F_1\}&=& \langle \partial_I N,\partial_\theta F_1\rangle
+ \langle \nabla_zN, { J}\nabla_zF_1\rangle \nonumber\\
&=& \langle \partial_I N,\partial_\theta F_1\rangle
+\sum_{i} \langle \nabla_{z_i}N, {{\rm i} J_{d_i}}\nabla_{z_i}F_1\rangle
\nonumber\\
&=&{\rm i}\sum_{k,i}
( \langle \;\langle k, \omega\rangle F^k_i,z_i\rangle +\langle A_iz_i, J_{d_i}F^k_i\rangle ) e^{{\rm i}\langle k, \theta\rangle}\;
\nonumber\\
&=&{\rm i}\sum_{k,i} \langle (\langle k, \omega\rangle I_{d_i}+A_iJ_{d_i})F^k_i, z_i\rangle
e^{{\rm i}\langle k, \theta\rangle}\; .
\end{eqnarray}
It follows that
$F^k_i$ are determined by the
linear algebraic system
$${\rm i}(\langle k,\omega\rangle { I_{d_i}}+ { A_iJ_{d_i} })F^k_i+ R^k_i=0,
\quad i\in {\Bbb N}, k\in {\Bbb Z}^d$$
\smallskip\noindent
Similarly, from
\begin{eqnarray}
\{N, F_2\}
&=& \langle \partial_IN, \partial_\theta F_2\rangle
+\sum_{i} \langle \nabla_{z_i}N, { {\rm i}J_{d_i}}\nabla_{z_i}F_2\rangle \nonumber\\
&=&{\rm i}\sum_{ |k|+|i-j|\ne 0} (\langle \;\langle k,\omega\rangle F^k_{ji}z_i,z_j\rangle
+\langle A_iz_i, J_{d_i}(F^k_{ji})^Tz_j\rangle +\langle A_jz_j, J_{d_j}F^k_{ji}z_i\rangle )
e^{{\rm i}\langle k, \theta\rangle}\; \nonumber\\
&=&{\rm i}\sum_{ |k|+|i-j|\ne 0} (\langle \;\langle k, \omega\rangle { F}^k_{ji}z_i,z_j\rangle
+\langle (A_jJ_{d_j}F^k_{ji}-F^k_{ji}J_{d_i}A_i)z_i,z_j\rangle )e^{{\rm i}\langle k, \theta\rangle}\; \nonumber\\
&=&{\rm i}\sum_{ |k|+|i-j|\ne 0}
\langle (\langle k,\omega\rangle F^k_{ji}+A_{j}J_{d_j}F^k_{ji}-
F^k_{ji}J_{d_i}A_i)z_i,z_j\rangle
e^{{\rm i}\langle k, \theta\rangle}\;
\end{eqnarray}
it follows that, $F^k_{ji}$ is determined by the following
matrix equation
\begin{equation}\label{matrixeq}
(\langle k,\omega\rangle I_{d_j}+ A_jJ_{d_j})F^k_{ji}- F^k_{ji}(J_{d_i}A_i)
={\rm i}R^k_{ji}, \quad |k|+
|i-j|\ne 0
\end{equation}
where $ F^k_{ji}, R^k_{ji}$ are $d_j\times d_i$ matrices defined in
(\ref{F}) and (\ref{R}). Exchanging $i, j$ we get the third equation in
(\ref{Feq}).
\nolinebreak\hfill\rule{2mm}{2mm
\medskip\noindent
The first two equations in (\ref{Feq}) are immediately solved in view of
(\ref{ODC}).
In order to solve the third equation in (\ref{Feq}), we need the following
elementary algebraic result from matrix theory.
\begin{Lemma}\label{matrixsolution}
Let $A, B, C$ be respectively $n\times n, m\times m, n\times m$ matrices,
and let $X$ be an $n\times m$ unknown matrix.
The matrix equation
\begin{equation} \label{meq} AX-XB=C,\end{equation}
is solvable if and only if $I_m\otimes A-B\otimes I_n$ is nonsingular. Moreover,
\[ \|X\|\le \|( I_m\otimes A-B\otimes I_n )^{-1}\|\cdot \|C\|.\]
\end{Lemma}
\smallskip\noindent
In fact, the matrix equation (\ref{meq}) is equivalent to the
(bigger) vector equation given by
$(I\otimes A-B\otimes I)X'=C'$ where $X', C'$ are vectors whose elements
are just the list (row by row) of the entries of $X$ and $C$.
For a detailed proof we refer the reader to the Appendix in \cite{Y} or
\cite{Lan}, p. 256.
\smallskip\noindent
\noindent
{\bf Remark}. Taking the transpose of the third equation in (\ref{Feq}),
one sees that $(F^k_{ij})^T$ satisfies the same equation of $F^k_{ji}$. Then
(by the uniqueness of the solution) it follows that
$F^k_{ji}=(F^k_{ij})^T$.
\medskip\noindent
\subsection{Estimates on the coordinate transformation}
\smallskip\noindent
We proceed to estimate $X_F$ and $\Phi_F^1$. We start with the following
\begin{Lemma}\label{Festimate}
Let $D_i=D(\frac i4 s, r_++\frac{i}4 (r-r_+))$,
$0 <i \le 4$. Then
\begin{equation}
\|X_F\|_{D_3, \Cal O} \;<\;{\rm c}\; \gamma^{-{\rm c}}\Gamma(r-r_+) \epsilon.
\end{equation}
\end{Lemma}
\Proof
By (\ref{ODC}), Lemma \ref{Fdef} and Lemmata \ref{extension1},
{\ref{extension2} in the Appendix, we have
\begin{eqnarray}
|F_{kl0}|_{\Cal O}&\le & |\langle k, \omega\rangle |^{-1} |P_{kl}|
\;<\;{\rm c}\; \gamma^{-{\rm c}}|k|^{\rm c}
e^{-|k|(r-r_+)}\epsilon
s^{2-2|l|}, \quad k\ne 0,\nonumber\\
\|F^k_i\|_{\Cal O}&=& \|(\langle k, \omega\rangle {I_{d_i}} + {A_iJ_{d_i}})^{-1}R^k_i\|\le
\|(\langle k, \omega\rangle I_{d_i}+ A_iJ_{d_i})^{-1}\|\cdot \|R^k_i\|\nonumber\\
&<& {\rm c}\ \gamma^{-{\rm c}}|k|^{\rm c} |R^k_i| ,\nonumber\\
\|F^k_{ij}\|_{\Cal O}&\le &
\|(\langle k, \omega\rangle {I_{d_id_j}} +(A_iJ_{d_i})\otimes I_{d_j}-
I_{d_i}\otimes (J_{d_j}A_j))^{-1}\|
\cdot \|R^k_{ij}\|\nonumber\\
&<& {\rm c}\ \gamma^{-{\rm c}}|k|^{\rm c}\|R_{ij}^k\|, \, \,|k|+|i-j|\ne 0.
\end{eqnarray}
Where $\|\cdot\|_{\Cal O}$ for matrix is similar to (\ref{Falpha}).
\smallskip\noindent
It follows that
\begin{eqnarray}
\frac 1{s^2}\|F_\theta\|_{D_2,\Cal O}
&\le & \frac 1{s^2}(\sum |f_{kl0}|\cdot|I^l|\cdot |k|\cdot |e^{{\rm i}\langle k, \theta\rangle}\; |
+\sum |F^k_i| \cdot|z_i|\cdot |k|\cdot |e^{{\rm i}\langle k, \theta\rangle}\; |\nonumber\\
&+&\sum \|F^k_{ij}\|\cdot |z_i|\cdot|z_j|\cdot |k|\cdot |e^{{\rm i}\langle k, \theta\rangle}\; |)
\nonumber\\
& \;<\;{\rm c}\; &
\gamma^{-{\rm c}}
\Gamma(r-r_+)\|X_R\|\nonumber\\
& \;<\;{\rm c}\; &\gamma^{-{\rm c}} \Gamma(r-r_+)\epsilon,
\end{eqnarray}
where $\Gamma(r-r_+)=\sup_{k} |k|^{\rm c} e^{-|k|\frac 14(r-r_+)}. $
\smallskip\noindent
Similarly,
\[ \|F_I\|_{D_2,\Cal O}=\sum_{|l|\le 1}|F_{kl0}|\cdot |e^{{\rm i}\langle k, \theta\rangle}\; |
\;<\;{\rm c}\; \gamma^{-{\rm c}} \Gamma(r-r_+)\epsilon.
\]
\smallskip\noindent
Now we estimate $\|X_{F^1}\|_{D_{2}, \Cal O}$.
Note that
\begin{eqnarray} \|F^1_{z_i}\|_{D_{2}, \Cal O}
&=&\|\sum_kF^k_i e^{-i<k,\theta>}\|_{D_{2}, \Cal O}
\nonumber\\
& \;<\;{\rm c}\; & \gamma^{-{\rm c}}\Gamma \sum_{k,i}|R^k_i|e^{|k|r}
\;<\;{\rm c}\; \gamma^{-{\rm c}}\Gamma \|\frac{\partial P_1}{\partial z_i}\|.
\end{eqnarray}
It follows that
\begin{eqnarray}
\|X_{F^1}\|_{D_{2}, \Cal O}& \;<\;{\rm c}\; &\sum_{i\in {\Bbb N}}
\|F^1_{z_i}\|_{D_{2}, \Cal O}i^ae^{i\rho}\nonumber\\
& \;<\;{\rm c}\; &\gamma^{-{\rm c}}\Gamma \sum_{i\in {\Bbb N}}
\|\frac{\partial P_1}{\partial z_i}\|i^ae^{i\rho}
\;<\;{\rm c}\; \gamma^{-{\rm c}}
\Gamma \epsilon,\nonumber
\end{eqnarray}
by the definition of the weighted norm.
\smallskip\noindent
Note that
\begin{eqnarray} \|F^2_{z_i}\|_{D_{2}, \Cal O}
&=&\|\sum_{k,j} (F^k_{ij}+(F^k_{ij})^T)z_j e^{{\rm i}\langle k, \theta\rangle}\; \|_{D_{2}, \Cal O}
\nonumber\\
& \;<\;{\rm c}\; & \gamma^{-{\rm c}}\Gamma \|\frac{\partial P_2}{\partial z_i}\|.
\end{eqnarray}
\smallskip\noindent
Similarly, we have
\begin{equation}\label{3.16}
\|X_{F^2}\|_{D_{2}, \Cal O} \;<\;{\rm c}\;
\gamma^{-{\rm c}} \Gamma \epsilon.
\end{equation}
The conclusion of the lemma follows from the above estimates.
\nolinebreak\hfill\rule{2mm}{2mm
\smallskip\noindent
In the next lemma, we give some estimates for $\phi_F^t$.
The following formula (\ref{3.32}) will be used to prove that our coordinate
transformations is well defined.
(\ref{3.33}) is for proving the convergence of the iteration.
\smallskip\noindent
\begin{Lemma}\label{(3.5)}
Let $\eta=\epsilon^{\frac 13}, D_{\frac i2\eta}=
D(r_++\frac {i-1}2(r-r_+),\frac{i}2 \eta s), i=1,2$.
We then have
\begin{equation}
\phi_F^t: D_{\frac 12\eta}
\to D_{\eta} ,\ \ \ 0\le t\le 1,
\label{3.32}
\end{equation}
if $\epsilon\ll (\frac 12\gamma^{-{\rm c}} \Gamma^{-1})^
{\frac 32}$. Moreover,
\begin{equation}
\|D\phi_F^1-Id\|_{D_{\frac 12\eta}} \;<\;{\rm c}\; \gamma^{-{\rm c}}\Gamma\epsilon.
\label{3.33}
\end{equation}
\end{Lemma}
\Proof
Let
$$\|D^mF\|_{D,\Cal O}
=\max \{ |\frac{\partial^{|i|+|l|+p}}{\partial \theta^{i}\partial I^{l}
\partial z^\alpha} F|_{D, \Cal O}, |i|+|l|+|\alpha|=m\ge 2\}.$$
\smallskip\noindent
Note that $F$ is polynomial in $I$ of order 1, in $z$ of order 2.
From\footnote{Recall the definition of the weighted norm in (\ref{weightednorm}).}
(\ref{3.16}) and the Cauchy inequality,
it follows
that
\begin{equation}\|D^mF\|_{D_1, \Cal O } \;<\;{\rm c}\;
\gamma^{-{\rm c}} \Gamma\epsilon,\label{3.17}
\end{equation}
for any $m\ge 2$.
\smallskip\noindent
To get the estimates for $\phi_F^t$, we start from the integral equation,
$$\phi_F^t=id+\int_0^tX_F\circ \phi_F^s\,ds$$
$\phi_F^t: D_{\frac 12\eta}
\to D_{\eta} ,\ \ \ 0\le t\le 1$, follows directly from (\ref{3.17}).
Since
$$D\phi_F^1=Id+\int_0^1(DX_F) D\phi_F^s\,ds=
Id+\int_0^1J(D^2F) D\phi_F^s\,ds,$$
it follows that
\begin{equation}\|D\phi_F^1-Id\|\le 2\|D^2F\| \;<\;{\rm c}\;
\gamma^{-{\rm c}}\Gamma \epsilon.
\label{3.36}
\end{equation}
The estimates of second order derivative $D^2\phi_F^1$ follows from
(\ref{3.17}).
\nolinebreak\hfill\rule{2mm}{2mm
\medskip\noindent
\subsection{Estimates for the new normal form}
\smallskip\noindent
The map $\phi_F^1$ defined above transforms $H$ into
$H_+=N_+ + P_+$(see (\ref{2.3}) and (\ref{hom}))
with
\begin{equation}
N_+= e_++\langle \omega_+, y\rangle +\frac 12\sum_{i\in {\Bbb Z}_+}\langle A^+_iz_i,z_i\rangle
\end{equation}
where
\begin{equation} e_+=e+P_{000},\
\omega_+=\omega+P_{0l0} (|l|=1),\
A_i^+=A_i+2 R^0_{ii}.
\end{equation}
\medskip\noindent
Now we prove that $N_+$ shares the same properties with $N$.
By the regularity of $X_P$ and by Cauchy estimates, we have
\begin{equation}\label{Rest}
|\omega_{+}-\omega|<\epsilon, \quad
\|R^{0}_{ii}\|<\epsilon i^{-\delta}
\end{equation}
with $\delta=\bar a -a>0 $.
It follows that
\begin{eqnarray} & & \|(A_i^+)^{-1}\|\le \frac
{\|A_i^{-1}\|}{1-2\|A_i^{-1}R^0_{ii}\|} \le 2\|A_i^{-1}\|,
\nonumber\\
& & \|(\langle k,\omega+P_{0l00}\rangle I_{d_i}-{J_{d_i}A_i^+})^{-1}\|
\le
\frac{\|(\langle k, \omega\rangle I_{d_i}+{A_iJ_{d_i}})^{-1}\|}
{1- \|(\langle k, \omega\rangle I_{d_i}+{A_iJ_{d_i}})^{-1}\|\epsilon}
\le (\frac{ |k|^\tau}{ \gamma_+})^{\bar d},
\label{3.2+}
\end{eqnarray}
provided $|k|^{\bar d\tau}\epsilon \;<\;{\rm c}\; (\gamma^{\bar d}-\gamma_+^{\bar d}) $.
\smallskip\noindent
Similarly, we have
\begin{equation}\label{4.30}
\|(\langle k,\omega+P_{0l00}\rangle I_{d_id_j}+
(A_i^+ J_{d_i})\otimes I_{d_j}- I_{d_i}\otimes
(J_{d_j}A_j^+))^{-1}\|\le
(\frac {|k|^\tau}{\gamma_+})^{\bar d^2},
\end{equation}
provided $|k|^{\bar d^2\tau}\epsilon \;<\;{\rm c}\; (\gamma^{\bar d^2}-\gamma_+^{\bar d^2}) $.
This means that in the next KAM step, small denominator conditions are
automatically satisfied for
$|k|<K$ where $K^{\bar d^2\tau}\epsilon \;<\;{\rm c}\; (\gamma^{\bar d^2}-\gamma_+^{4\bar
d^2}) $.
\smallskip\noindent
The following bounds wil be used later for the measure estimates.
\begin{equation}\label{a+}
|\frac{\partial^l(\omega_+-\omega)}
{ \partial\xi^l}|_{\Cal O}\le \epsilon,\quad
|\frac{\partial^l(A_{i}^+-A_{i})}
{ \partial\xi^l}|
_{\Cal O} \;<\;{\rm c}\; \epsilon i^{-\delta},\end{equation}
for $|l|\le \bar d^2$ by the definition of the norm.
\subsection{ Estimates for the new perturbation}
To complete the KAM step we have to estimate the new error term.
\smallskip\noindent
By the definition of $\phi^1_F$ and
Lemma \ref{(3.5)},
$$ H\circ \phi^1_F
=N_++ P_+
$$
is well defined in $D_{\frac 12\eta}.$
Moreover, we have the following estimates
\begin{eqnarray}
\|X_{P_+}\|_{D_{\frac 12\eta}}
&=& \|X_{\int_0^1dt\int_0^s\{\{N+R, F\},F\}\circ \phi^s_F
+\{R,F\}+(P-R)\circ\phi^1_F}\|_{D_{\frac 12\eta}}
\nonumber \\
&\le&\|X_{(\int_0^1dt\int_0^t\{\{N+R, F\},F\}\circ
\phi^s_F }\|_{D_{\frac 12\eta}}
+\|X_{(P-R)\circ\phi^1_F}\|_{D_{\frac 12\eta}}
\nonumber \\
&\le &\|X_{\{\{N+R, F\},F\}}\|_{D_{\eta}}
+\|X_{P-R}\|_{D_{\eta}}\nonumber\\
& \;<\;{\rm c}\; & \gamma^{-{\rm c}} \Gamma^2
\epsilon^{\frac 43} \;<\;{\rm c}\; \epsilon_+
\label{3.37}
\end{eqnarray}
by (\ref{P-R}) and Lemma \ref{FG}.
\smallskip\noindent
Thus, there exists a big constant $c$,
independent of iteration steps, such that
\begin{equation}\label{P+}
\|X_{P_+}\|_{r_+, s_+}=\|X_{P_+}\|_{D_{\frac 12\eta}}^{\bar a, \rho}
\le c \gamma^{-{\rm c}}\Gamma^2\eta\epsilon=c\epsilon_+.
\end{equation}
\smallskip\noindent
The KAM step is now completed.
\section{Iteration Lemma and Convergence}
For any given $s, \epsilon, r, \gamma$, we define, for all $\nu\ge 1$, the
following sequences
\begin{eqnarray}\label{iterationconstants}
r_{\nu} &=&r(1- \sum_{i=2}^{\nu+1 }2^{-i}),\nonumber\\
\epsilon_{\nu} &=& c\gamma_{\nu}^{-\rm c}\Gamma(r_{\nu-1}-r_{\nu})^2
\epsilon_{\nu-1}^{\frac 43}.\nonumber\\
\gamma_{\nu} &=&\gamma(1-\sum_{i=2}^{\nu+1 }2^{-i})\nonumber\\
\eta_{\nu}& =& \frac 12\epsilon_{\nu}^{\frac 13},\quad
L_{\nu} = L_{\nu-1}+\epsilon_{\nu-1}, \nonumber\\
s_{\nu} &=& \frac 12\eta_{\nu-1} s_{\nu-1}= 2^{-\nu}
(\prod_{i=0}^{\nu-1}\epsilon_i)^{\frac 13}s_0,\nonumber\\
K_{\nu} &=& \frac c2\left(\epsilon_{\nu-1}^{-1}
(\gamma_{\nu-1}^{\bar d^2}
-\gamma_{\nu}^{\bar d^2})\right)^{\frac1{\bar d^2\tau}},\nonumber\\
D_{\nu} &=& D_{a,\rho}(r_{\nu}, s_{\nu}),
\end{eqnarray}
where $c$ is the constant in (\ref{P+}). The parameters
$r_0, \epsilon_0, \gamma_0, L_0, s_0, K_0$ are defined respectively to be
$r, \epsilon, \gamma, L, s, 1$.
\medskip\noindent
Note that
$$
\Psi(r) =\prod_{i=1}^\infty [\Gamma(r_{i-1}-r_i)]^{2(\frac 34)^{i}},$$
is a well defined finite function of $r$.
\medskip\noindent
\subsection{Iteration Lemma}
\smallskip\noindent
The analysis of the preceeeding section can be summarized as follows.
\begin
{Lemma}\label{iteration} Suppose that $\epsilon_0=\epsilon(d, \bar
d, \delta, \delta_1,
\bar a-a, L, \tau, \gamma)$
is small enough.
Then the following holds for all $\nu\ge 0$.
Let
$$N_\nu=e_\nu+\langle \omega_\nu(\xi),I\rangle
+\sum_{i\in {\Bbb N}}\langle A^\nu_i(\xi) z_i, z_i\rangle
,$$
be a normal form with parameters $\xi$ satisfying
\begin{eqnarray}
& &|\langle k,\omega_\nu\rangle ^{-1} |<\frac{|k|^{\tau}}{\gamma_\nu }, \;\;\;
\|({\rm i}\langle k,\omega_\nu\rangle { I_{d_i}}+{ A^\nu_iJ_{d_i} })^{-1}\|<
(\frac {|k|^{\tau}}{\gamma_\nu })^{\bar d}
\nonumber\\
& &\|({\rm i}\langle k,\omega_\nu\rangle I_{d_id_j}+(A_i^\nu J_{d_i} )\otimes I_{d_j}
-I_{d_i}\otimes (J_{d_j}A_j^\nu )
)^{-1}\|
<(\frac{|k|^{\tau}}{\gamma_\nu })^{\bar d^2}
\end{eqnarray}
on a closed set $\Cal O_\nu$ of $R^n$ for
all $k\ne 0, i, j\in {\Bbb Z}$.
Moreover,
suppose that $\omega_\nu(\xi), A_{i}^\nu(\xi)$ are $C^{\bar d^2}$
smooth and satisfy
$$|\frac{\partial^{\bar d^2}(\omega_\nu-\omega_{\nu-1})}
{\partial\xi^{\bar d^2}}|
\le \epsilon_{\nu-1},\
|\frac{\partial^{\bar d^2}(A_{i}^\nu-A_{i}^{\nu-1})}
{\partial\xi^{\bar d^2}}
|\le \epsilon_{\nu-1} i^{-\delta},$$
on $\Cal O_\nu$ in Whitney's sense.
\smallskip\noindent
Finally, assume that $$\|X_{P_\nu}\|_{D_\nu,\Cal O_{\nu}}^{\bar a,\rho}
\le \epsilon_{\nu}.$$
Then, there
is a subset $\Cal O_{\nu+1}\subset \Cal O_{\nu}$,
$$\Cal O_{\nu+1} =\Cal O_{\nu}-
\cup_{ |k|\ge K_{\nu+1}} {\Cal R}^{\nu+1} _{kij}(\gamma_{\nu})$$
where
\begin{displaymath}
\Cal R_{kij}^{\nu+1}(\gamma_{\nu+1})=
\left\{\xi\in \Cal O_\nu:
|\; \stackrel{\scriptstyle |\langle k,\omega_{\nu+1}\rangle ^{-1}|>\frac
{|k|^{\tau}}{\gamma_{\nu}},\quad
\|(\langle k,\omega_{\nu}\rangle I_{2m}+
(A_i^{\nu+1}J_{d_i})^{-1} \|\ge (\frac
{|k|^{\tau}}{\gamma_{\nu}})^{\bar d},\, \mbox{or}}
{\scriptstyle \|(\langle k,\omega_{\nu+1}>I_{d_id_j}+(A_j^{\nu+1}J_{d_i})
\otimes I_{d_j}-I_{d_i}\otimes
(J_{d_j}A_j^{\nu+1} ))^{-1}\|
> (\frac
{|k|^{\tau}}{\gamma_{\nu}}
})^{\bar d^2}\;\right\},
\end{displaymath}
with $\omega_{\nu+1}= \omega_\nu+P_{0l0}^\nu$,
and a symplectic change of variables
\begin{equation}
\Phi_\nu: D_{\nu+1}\times \Cal O_{\nu+1} \to D_\nu,
\label{4.1}
\end{equation}
such that
$H_{\nu+1}=H_\nu \circ\Phi_\nu$, defined on
$D_{\nu+1}\times \Cal O_{\nu+1} $,
has the form
\begin{equation}H_{\nu+1}=e_{\nu+1}+\langle \omega_{\nu+1},I\rangle
+\sum_{i\in {\Bbb N}} \langle A_{i}^{\nu+1}z_i, z_i\rangle +P_{\nu+1},
\label{4.2}
\end{equation}
satisfying
\begin{equation}\label{3.6}
\max_{l\le \bar d^2}|\frac{\partial^l(\omega_{\nu+1}(\xi)-\omega_\nu(\xi))}
{\partial\xi^l}|\le \epsilon_{\nu}, \;\;
{\rm max}_{|l|\le \bar d^2}|\frac{\partial^{l}(A_{i}^{\nu+1}(\xi)-
A_{i}^\nu)}{\partial\xi^{l}}| \le \epsilon_{\nu} i^{-\delta},
\end{equation}
\begin{equation}
\|X_{P_{\nu+1}}\|_{D_{\nu+1},\Cal O_{\nu+1}}^{\bar a,\rho}
\le \epsilon_{\nu+1}.
\label{4.3}\end{equation}
\end{Lemma}
\subsection{ Convergence}
Suppose that the assumptions of Theorem 1 are satisfied. To apply the iteration lemma with $\nu=0$, recall that
$$\epsilon_0=\epsilon, \gamma_0=\gamma, s_0=s,L_0=L, N_0=N, P_0=P,$$
\begin{displaymath}
\Cal O_0=
\left\{\xi\in \Cal O:
|\; \stackrel{\scriptstyle |\langle k,\omega\rangle ^{-1}|<\frac
{|k|^{\tau}}{\gamma },
\|(\langle k,\omega\rangle I_{d_i}+A_iJ_{d_i} )^{-1}\| < (\frac
{|k|^\tau}\gamma )^{\bar d},\, \mbox{or}}
{\scriptstyle \|(\langle k,\omega\rangle I_{d_id_j}+(A_iJ_{d_i})\otimes I_{d_j}
-I_{d_i}\otimes
(J_{d_j}A_j ))^{-1}\| < (\frac {|k|^{\tau}}
\gamma)^{\bar d^2}}\;\right\},
\end{displaymath}
(with $\epsilon$ and $\gamma$ small enough).
Inductively, we obtain
the following sequences
$$\Cal O_{\nu+1}\subset \Cal O_{\nu},$$
$$\Psi^\nu=\Phi_1\circ\cdots\circ\Phi_\nu:
D_{\nu+1}\times \Cal O_{\nu} \to D_0, \nu\ge 0,$$
$$H\circ\Psi^\nu=H_{\nu+1}=N_{\nu+1}+P_{\nu+1}.$$
\smallskip\noindent
Let $\Cal O_{\gamma}=\cap_{\nu=0}^\infty\Cal O_\nu$.
As in \cite{P2}, thanks to Lemma \ref{(3.5)},
we may conclude that $N_\nu, \Psi^\nu, D\Psi^\nu, \omega_{\nu+1}$ converge uniformly
on $D_\infty\times \Cal O_\gamma=D(\frac 12 r, 0,0)\times
\Cal O_\gamma$ with
$$N_\infty=e_\infty+\langle \omega_\infty,I\rangle +\langle A_\infty z,z\rangle
=e_\infty+\langle \omega_\infty,I\rangle
+\sum_{i\in {\Bbb N}}\langle A_i^\infty z_i, z_i\rangle
,$$
Since
$$\epsilon_{\nu+1} =
c\gamma_{\nu}^{-\rm c}\Gamma(r_{\nu}-r_{\nu+1})
\epsilon_{\nu}\le
(c\gamma^{-\rm c}\Psi(r)\epsilon)^{(\frac 43)^\nu}.$$
It follows that $ \epsilon_{\nu+1}\to 0$ provided
$\epsilon$ is sufficiently small.
\smallskip\noindent
Let $\phi^t_H$ be the flow of $X_H$. Since
$H\circ\Psi^\nu=H_{\nu+1}$, we have that
\begin{equation}\phi^t_H
\circ \Psi^\nu= \Psi^\nu \circ \phi^t_{H_{\nu+1}}.
\label{4.10}
\end{equation}
The convergence of $\Psi^\nu,D\Psi^\nu,\omega_{\nu+1}, X_{H_{\nu+1}}$ implies that
one can take limit in (\ref{4.10}) so as to get
\begin{equation}\label{4.11}
\phi^t_H\circ \Psi^\infty= \Psi^\infty\circ\phi^t_{H_\infty},
\end{equation}
on $D(\frac 12r, 0,0)\times \Cal O_\gamma$, with
$$
\Psi^\infty:D(\frac 12r,0,0)\times\Cal O_\gamma
\to \Cal P_{a, \rho}\times {\Bbb R}^d.
$$
\smallskip\noindent
From (\ref{4.11}) it follows that
$$\phi^t_H(\Psi^\infty({\Bbb T}^d\times \{
\xi\}))=\Psi^\infty\phi^t_{N_\infty}
({\Bbb T}^d\times \{\xi\})=\Psi^\infty ({\Bbb T}^d\times \{\xi\}),$$
for $\xi\in \Cal O_\gamma$.
This means that
$\Psi^{\infty} ({\Bbb T}^d\times \{\omega\})$ is an embedded torus invariant
for the original perturbed Hamiltonian
system at $\xi\in \Cal O_\gamma.$
We remark here the frequencies $\omega_\infty(\xi)$
associated to $\Psi^{\infty} ({\Bbb T}^d\times\{\xi\})$
is slightly different from $\xi$.
The normal behaviour of the invariant torus
is governed by the matrix $A^{\infty}_i=\sum_{\nu\in {\Bbb N}}A^\nu_i$.
\nolinebreak\hfill\rule{2mm}{2mm
\section{ Measure Estimates}\label{measureestimate}
At each KAM step, we have to exclude the following resonant
set of $\xi$'s: $$\Cal R^{\nu}=\bigcup_{|k|>K_{\nu},i,j}(\Cal R_k^{\nu} \cup\Cal
R_{ki}^{\nu}
\cup\Cal R_{kij}^{\nu})\ ,$$ the sets $ \Cal R_k^{\nu}, \Cal R_{ki}^{\nu},
\Cal R_{kij}^{\nu}$ being respectively
\begin{eqnarray}
& &\{\xi\in \Cal O_\nu: |\langle k,\omega_{\nu}\rangle ^{-1}|>\frac
{|k|^{\tau}}{\gamma_{\nu}}\},\quad
\{\xi\in \Cal O_\nu:\,\, \|\Cal M_1^{-1}\|> (\frac
{|k|^{\tau}}{\gamma_{\nu}})^{\bar d}\},\nonumber\\
& & \mbox{and}\;\; \{\omega\in \Cal O_\nu:\,\, \|\Cal M_2^{-1}\|> (\frac
{|k|^{\tau}}{\gamma_{\nu}})^{\bar d^2}\},
\end{eqnarray}
where
\begin{eqnarray}
&\Cal M_1 &= \langle k,\omega_{\nu}\rangle I_{d_i}+A_i^{\nu}J_{d_i}\nonumber\\
&\Cal M_2 & = \langle k,\omega_{\nu}\rangle I_{d_id_j}+
(A_j^{\nu}J_{d_j})\otimes I_{d_i}-I_{d_j}\otimes
(J_{d_i}A_i^{\nu} )
.\label{kij}
\end{eqnarray}
We include in the
set $\{\xi\in \Cal O: \|M(\omega)^{-1}\|>C\} $
also the $\xi$'s for which $M$ is not invertible.
Remind that $\omega_{\nu}(\xi)=\xi+
\sum_{j=0}^{\nu-1} P_{000}^j(\xi)$ with\footnote{Recall (\ref{Rest}),
(\ref{3.6}).
}
$|\sum P_{000}^j(\xi)|_{C^{\bar d^2}}\le \epsilon$,
$A_i^{\nu}= A_i+2\sum_\nu R^{0,\nu}_{ii}$ with $\|\sum_\nu R^{0,\nu}_{ii}\|
=O(\epsilon i^{-\delta})$.
\begin{Lemma}\label{Rkij} There is a constant $K_0$ such that,
for any $i,j,$ and $ |k|>K_0$,
\[{\rm meas }(\Cal R_{k}^{\nu}\cup\Cal R_{ki}^{\nu}
\cup\Cal R_{kij}^{\nu}) \;<\;{\rm c}\;
\frac{\gamma}{ |k|^{\tau-1}}.\]
\end{Lemma}
\Proof
As it is well known
\[{\rm meas}\,(\Cal R_{k}^{\nu})\le \frac{\gamma_{\nu}}{|k|^\tau}.\]
The set $\Cal R_{ki}^{\nu}$ is empty
if $i>{\rm const}\;|k|$, while, if $i\le {\rm const}\;|k|,$
from Lemmata \ref{3}, \ref{A.1} there follows that
\[{\rm meas }(\Cal R_{ki}^{\nu}) \;<\;{\rm c}\;
\frac{\gamma_{\nu}}{ |k|^{\tau-1}}.\]
\smallskip\noindent
We now give a detailed proof for the most complicated estimate, i.e., the estimate
on the measure of the set
$\Cal R_{kij}^{\nu}$. Rewrite $\Cal M_2$ as
\[\Cal M_2
\equiv \Cal A_{ij}+\Cal B_{ij}^{\nu},
\] with
\begin{equation}\label{calaij}
\Cal A_{ij}= \langle k,\omega_{\nu+1}\rangle I_{d_id_j}+\lambda_j\,{\rm Diag}
(I_{d_j/2}, -I_{d_j/2})\otimes I_{d_i}
-\lambda_i I_{d_j}\otimes
{\rm Diag}\;(-I_{d_i/2}, I_{d_i/2}).
\end{equation}
The matrix $\Cal A_{ij}$ is diagonal with entries
$\lambda_{kij}=\langle k,\omega_{\nu}\rangle \pm \lambda_i\pm \lambda_j$
in the diagonal where $\lambda_i, \lambda_j$ are given in (\ref{asymp1}) and
$\pm$ sign depends on the position.
$\Cal B_{ij}^{\nu}$ is a matrix of size $O(i^{-\delta}+
j^{-\delta})$ since $ A_i^{\nu}=
A_i+ B_i+O(i^{-\delta})=A_i+O(i^{-\delta}) $ by (\ref{asymp2})
and (\ref{Rest}).
\medskip\noindent
In the rest of the proof we drop in the notation the indices $i,j$ since they
are fixed.
Now either all $\lambda_{kij}\le|k|$ or there are some diagonal elements
$\lambda_{kij}> |k|$. We first consider the latter case.
By permuting rows and columns, we can find two non-singular
matrices $Q_1, Q_2$ with elements 1 or 0 such that
\begin{equation}\label{7.3}
Q_1(\Cal A +\Cal B^{\nu})Q_2
=\left(\begin{array}{cc}
A_{11} & 0\\
0 & A_{22}
\end{array}\right)+
\left(\begin{array}{cc}
\tilde B_{11} & \tilde B_{12}\\
\tilde B_{21} & \tilde B_{22}
\end{array}\right)
\end{equation}
where $A_{11}, A_{22}$ are diagonal matrices and $A_{11}$
contains all diagonal elements $\lambda_{kij}$ which are bigger than
$|k|$. Moreover, defining $Q_3, Q_4, D$ as
\[Q_3=\left(\begin{array}{cc}
I & \tilde 0\\
-\tilde B_{21}( A_{11}+\tilde B_{11})^{-1}
& I
\end{array}\right)
,\quad
Q_4=\left(\begin{array}{cc}
I & -( A_{11}+\tilde B_{11})^{-1}\tilde B_{12}\\
0 & I
\end{array} \right ),\]and
\begin{equation}
\label{D}
D=A_{22}+ \tilde B_{22}-\tilde B_{21}(A_{11}+\tilde B_{11})^{-1}
\tilde B_{12}= A_{22}+O(i^{-\delta}+j^{-\delta}),
\end{equation}
we have
\begin{equation} Q_3Q_1\left(\Cal A +\Cal B^{\nu+1}\right)Q_2Q_4=
\left(\begin{array}{cc}
A_{11}+B_{11} & 0\\
0 & D
\end{array}\right)
\end{equation}
For $\xi\in \Cal O$ such that $D$ is invertible,
we have
\begin{equation}\label{inverse}
(\Cal A+\Cal B^{\nu})^{-1}=Q_2Q_4\left(\begin{array}{cc}
( A_{11}+B_{11})^{-1} & 0\\
0 & D^{-1}
\end{array}\right)Q_3Q_1.
\end{equation}
Since the norm of $Q_1, Q_2, Q_3, Q_4, (A_{11}+B_{11})^{-1}$ are uniformly
bounded,
it follows from (\ref{inverse}) that
\begin{equation}\label{7.6}
\{\xi\in \Cal O_\nu:
\|(\Cal A+\Cal B^{\nu})^{-1}\|
> (\frac
{|k|^{\tau}}{\gamma_{\nu}})^{\bar d^2}\}\nonumber\\
\subset
\{\xi\in \Cal O_{\nu}:
\|D^{-1}\| \;>{\rm c}\; (\frac
{|k|^{\tau}}{\gamma_{\nu}})^{\bar d^2}\}.
\end{equation}
If all $\lambda_{kij} \;<\;{\rm c}\; |k|$ we simply take $D=\Cal A+{\Cal B}^{\nu}$.
Since all elements in
$D$ are of size $O(|k|)$, by Lemma \ref{3} in the
Appendix, we have
\begin{equation}\label{7.6'}
\{\xi\in \Cal O_{\nu}:
\|D^{-1}\|\;>{\rm c}\; \;( \frac
{|k|^{\tau}}{\gamma_{\nu}})^{\bar d^2}\}
\subset
\{\xi\in \Cal O_{\nu}:
|\det D| \;<\;{\rm c}\; (\frac{\gamma_{\nu}}
{|k|^{\tau-1}})^{\bar d^2}\}.
\end{equation}
\smallskip\noindent
Let $N$ denote the dimension of $D$ (which is not bigger than
$ \bar d^2$).
Since $D=A_{22}+O(i^{-\delta}+j^{-\delta})$,
the $N^{\rm th}$ order derivative of $\det D$ with respective to some
$\xi_i$ is bounded away from zero by
$\frac 1{2d} |k|^{N}$ (provided $|k|$ is bigger enough).
From (\ref{7.6}), (\ref{7.6'}) and Lemma \ref{A.1}, it follows that
\begin{eqnarray}\label{7.6''}
{\rm meas}\,\Cal R_{kij}^{\nu}&=&
{\rm meas} \,\{\xi\in \Cal O_\nu:
\|(\Cal A+\Cal B^{\nu})^{-1}\|
> (\frac
{|k|^{\tau}}{\gamma_{\nu}})^{\bar d^2}\}
\nonumber\\
&\le & {\rm meas}\{\xi\in \Cal O_{\nu}:
|\det D| \;<\;{\rm c}\; (\frac{\gamma_{\nu}}
{|k|^{\tau-1}})^{\bar d^2}\}\nonumber\\
& \;<\;{\rm c}\; & (\frac{\gamma_{\nu}}{|k|^{\tau-1}})^{\frac {\bar d^2}N}
\;<\;{\rm c}\; \frac{\gamma}{|k|^{\tau-1}}.
\end{eqnarray}
This proofs the lemma.
\nolinebreak\hfill\rule{2mm}{2mm
\begin{Lemma}\label{empty}
If $i\;>{\rm c}\; |k|$, then $\Cal R^{\nu}_{ki}=\emptyset$;
If $\max\{i, j\}\;>{\rm c}\; |k|^{\frac 1{b-1}}, i\ne j$ for $b>1$ or
$|i-j|>{\rm const}\; |k|$ for $b=1$, then $\Cal R^{\nu}_{kij}=\emptyset$
where the constant {\rm c} depends on the diameter of $\Cal O$.
\end{Lemma}
\Proof As above, we only consider the most complicated case, i.e., the case of $\Cal
R^{\nu}_{kij}$. Notice that $\max\{i, j\}> {\rm const}\;|k|^{\frac 1{b-1}}$ for
$b>1$ or
$|i-j|>{\rm const}\; |k|$ for $b=1$ implies
\begin{eqnarray}
|\lambda_i\pm \lambda_j|&=&
(j^b-i^b)(1+O(i^{-\delta}+j^{-\delta}))
\nonumber\\
&\ge & \frac 12 |j-i|(i^{b-1}+j^{b-1})(1+O(i^{-\delta}+j^{-\delta}))
\ge {\rm const}\;|k|.\end{eqnarray}
It follows that $\Cal A_{ij}$ defined in (\ref{calaij}) is invertible and
\[\|( \Cal A_{ij})^{-1}\|<|k|^{-1}.\]
By Neumann series,
we have $\|(\Cal A_{ij}+\Cal B_{ij}^{\nu})^{-1}\|< 2|k|^{-1}$ for large $k$ (say
$|k|>K_0$),
i.e, $\Cal R^{\nu}_{kij}=\emptyset$.
\nolinebreak\hfill\rule{2mm}{2mm
\begin{Lemma}
For $b\ge 1$, we have
\[{\rm meas}(\bigcup_{\nu\ge 0}\Cal R^{\nu})
={\rm meas}\;\bigcup_{\nu, |k|>K_{\nu},i,j}(\Cal R_k^{\nu} \cup\Cal R_{ki}^{\nu}
\cup\Cal R_{kij}^{\nu}) \;<\;{\rm c}\; \gamma^{\frac
\delta{1+\delta}}. \]
\end{Lemma}
\Proof The measure estimates for $\Cal R^0$
comes from our assumption (\ref{ndc}). We then consider the estimate
\[{\rm meas}(\bigcup_\nu\bigcup_{|k|>K_{\nu}}\bigcup_{i,j}\Cal R^{\nu}_{kij}) \ ,\]
which is the most complicate one.
\smallskip\noindent
Let us consider separately the case $b>1$ and the case $b=1$.
We first consider $b>1$.
By Lemmata \ref{Rkij}, \ref{empty},
if $|k|>K_0$ and $i\ne j$, we have
\begin{equation}\label{inej}
{\rm meas}\;(\bigcup_{i\ne j}\Cal R_{kij})=
{\rm meas}\;(\bigcup_{i\ne j; i,j<C|k|^{\frac 1{b-1}}}\Cal R^k_{ij})
\;<\;{\rm c}\; \frac{|k|^{\frac 2{b-1}}
\gamma }{|k|^{ \tau-1}}\;>{\rm c}\;
\frac\gamma {|k|^{ \tau-1-\frac 2{b-1}}}. \end{equation}
For $i=j$.
As in Lemma \ref{Rkij}, we can find $Q_1, Q_2$ so that (\ref{7.3}) holds with
the diagonal elements of $A_{11}$ being
$<k, \omega_{\nu}>\pm 2\lambda_i$ and $A_{22}= <k, \omega_{\nu}>I$.
Repeating the arguments in Lemma \ref{Rkij}, we get (\ref{7.6'}) and
\begin{eqnarray}\label{7.13}
\Cal R_{kii}^{\nu} &\subset&
\{\xi:
|\det D| \;<\;{\rm c}\;
(\frac{\gamma_{\nu}}
{|k|^{\tau-1}})^{\bar d^2}\}\nonumber\\
&=&
\{\xi:\prod |\langle k,\omega_{\nu}\rangle +O(i^{-\delta})|
\;<\;{\rm c}\;
(\frac{\gamma_{\nu}}
{|k|^{\tau-1}})^{\bar d^2}\}\nonumber\\
&\subset&
\{\xi: |\langle k, \omega_{\nu}\rangle | \;<\;{\rm c}\; ( \frac \gamma{|k|^{\tau-1}}
+\frac 1{i^{\delta}})\}\equiv \Cal Q_{ii}^k.\end{eqnarray}
Since $\Cal Q_{ii}^k\subset \Cal Q_{i_0i_0}^k $ for $i\ge i_0$, using
(\ref{7.6''}), we find
that
\[ {\rm meas}\;(\bigcup_i \Cal R_{kii})
\le \sum_{i<i_0}|\Cal R_{kii} |+|\Cal Q_{i_0i_0}^k|
\;<\;{\rm c}\; (\frac{i_0\gamma} {|k|^{\tau-1}}+\frac 1{i_0^{-\delta}})\]
for any $i_0$.
Following P\"oschel (\cite{P2}), we choose
$i_0= (\frac {|k|^{\tau-1}}{\gamma})^\frac 1{1+\delta}$,
so that
\begin{equation}\label{7.13'}
{\rm meas }\;(\bigcup_i \Cal R_{kii}| \;<\;{\rm c}\;
(\frac {\gamma}{|k|^{\tau-1}})^ {\frac\delta{1+\delta}}\end{equation}
Let $\tau>\max\{d+2+\frac 2{b-1},
(d+1)^{\frac{1+\delta}\delta}+1\}$. As in (\ref{inej}), (\ref{7.13'}), we find
\begin{eqnarray}
& &
{\rm meas}\;(\bigcup_{|k|>K_\nu}\bigcup_{i, j}\Cal R_{kij}^{\nu}
(\gamma_{\nu}))
={\rm meas}\;(\bigcup_{|k|>K_\nu}\bigcup_{i\ne j}\Cal R_{kij}^{\nu}
(\gamma_{\nu}))\nonumber\\
\ \ \ & & +{\rm meas}\;(\cup_{|k|>K_\nu}\bigcup_{i }\Cal R_{kii}^{\nu}(\gamma_{\nu+1}))
\;<\;{\rm c}\; K_\nu^{-1}\gamma^{\frac\delta{1+\delta}}.\nonumber
\end{eqnarray}
The quantity
${\rm meas}(\bigcup_\nu\bigcup_{|k|>K_{\nu}}\bigcup_{i,j}
\Cal R^{\nu}_{kij})$ is then
bounded by
\begin{equation}
\sum_{\nu\ge 1} {\rm meas}(
\bigcup_{|k|>K_\nu}\bigcup_{i, j}\Cal R_{kij}^{\nu}(\gamma_{\nu}))
\;<\;{\rm c}\; \gamma^{\frac \delta{1+\delta}}\sum_{\nu\ge 0}K_{\nu}^{-1}
\;<\;{\rm c}\; \gamma^{\frac \delta{1+\delta}},
\end{equation}
provided $\tau>\max\{d+2+\frac 2{b-1},
(d+1)^{\frac{1+\delta}\delta}+1\}$. This concludes the proof for $b>1$.
\medskip\noindent
Consider now $b=1$. Without loss of generality, we assume $j\ge i$ and
$j=i+m$. Note that Lemma
{\ref{empty} implies $\Cal R^k_{ij}=\emptyset$ for $m>C|k|$.
Following the scheme of the above proof, we
find
\begin{eqnarray}
\bigcup_{k,i,j}\Cal R_{kij}&=&\bigcup_{k,i,m}\Cal R_{ki,i+m}=
\bigcup_{k, m<C|k|}\bigcup_i\Cal R_{ki,i+m}\nonumber\\
&\subset &\bigcup_{k, m<C|k|}(\bigcup_{i<i_0}\Cal R_{ki_0,i_0+m}\cup
\Cal Q_{ki_0,i_0+m} ).
\end{eqnarray}
where
\[\Cal Q_{k i_0, i_0+m}=\{\xi: |\langle k, \omega_{\nu}\rangle +m
| \;<\;{\rm c}\; ( \frac \gamma{|k|^{\tau-1}}
+\frac 1{i^{-\delta}})\}.
\]
Again, taking $i_0^{1+\delta}= \frac{|k|^{\tau-1}}{\gamma},$
we have, for fixed $k$,
\begin{eqnarray}
|\bigcup_{i,j}\Cal R_{ij}^k| & \;<\;{\rm c}\; & \sum_{m<C|k|}(\frac{i_0\gamma}{|k|^{\tau-1}}
+i_0^{-\delta})\nonumber\\
& \;<\;{\rm c}\; & |k|( \frac \gamma {|k|^{\tau-1}})^{\frac \delta{1+\delta}}\ .
\end{eqnarray}
As in the case $b>1$, we have that
${\rm meas}(\bigcup_\nu\bigcup_{|k|>K_{\nu}}\bigcup_{i,j}\Cal R^{\nu}_{kij})$
is bounded by
$O(\gamma^{\frac\delta {1+\delta}})$
if $\tau> (d+1)^{\frac {1+\delta}\delta} +1$.
\nolinebreak\hfill\rule{2mm}{2mm
\noindent
{\bf Remark} \ In (\ref{7.13}), $|\det D|=\prod|\langle k, \omega\rangle +O(i^{-\delta})|$
(guaranteed by the regularity property) is crucial for the proof. But it is
not necessary for the periodic solution case, i.e., $d=1$.
since $\Cal R_{k, i+m, i}^{\nu}
=\emptyset$ if, $ i\;>{\rm c}\; > |k|\ll 1$ are sufficiently large.
\section{Appendix}
\smallskip\noindent
{\it Proof of Proposition \ref{prop}} From the hypotheses there follows that the
eigenfuctions $\phi_n$ are analytic (respectively, smooth) and bounded with, in
particular,
$$\sup_{\Bbb R} (|\phi_n'|+|\phi_n''|)\le\ {\rm const}\ \mu_n\ .$$
Thus, the sum defining $u(t,x)$ is uniformly convergent in $I\times [0,2\pi]$. Since
$$
\frac{\partial G}{\partial q_n} = - \frac{1}{\sqrt{\lambda_n}} \int f(\sum_k
\frac{q_k}{\sqrt{\lambda_k}} \phi_k) \phi_n\ ,
$$
one has
$$
|q_n|\le {\rm const}\ \frac{e^{-n\rho}}{n^a}\ ,\quad
|\dot q_n|\le {\rm const}\ \lambda_n \frac{e^{-n\rho}}{n^a}
\le {\rm const}\ \frac{e^{-n\rho}}{n^{a-1}}\ ,
$$
$$|\ddot q_n|\le {\rm const}\ \frac{e^{-n\rho}}{n^{a+1}}\ .
$$
Thus (if $a$ is big enough, in the smooth case) $u(t,x)$ is a $C^2$ function and
\begin{eqnarray}
u_{tt}+Au &=& \sum \frac{\ddot q_n}{\sqrt{\lambda}} \phi_n +
\frac{q_n}{\sqrt{\lambda_n}} A \phi_n \nonumber\\
&=& \sum \Big( \int f(u) \phi_n\Big) \phi_n = f(u)\ ,\nonumber\\
\end{eqnarray}
where in the last equality we used the fact that $f(u)$ is a smooth periodic
function.
\nolinebreak\hfill\rule{2mm}{2mm
\begin{Lemma}
\[\|FG\|_{D(r,s)}\le \|F\|_{D(r,s)} \|G\|_{D(r,s)}.\]
\end{Lemma}
\Proof
Since $(FG)_{klp}=\sum_{l}F_{k-k',l-l',p-p'}G_{k'l'p'}$.
we have that
\begin{eqnarray}
\|FG\|_{D(r,s)}&=&\sup_D \sum_{klp}|(FG)_{klp}| \ |y|^l\ |z^\alpha|
e^{|k|r}
\nonumber\\
&\le &\sup_D\sum_{klp} \sum_{l'}|F_{k-k',l-l',p-p'}G_{k'l'p'}|\
|y|^l|z^\alpha| e^{|k|r}\nonumber\\
&=& \|F\|_{D(r,s)} \|G\|_{D_(r,s)}
\end{eqnarray}
and the proof is finished.
\nolinebreak\hfill\rule{2mm}{2mm
\begin{Lemma}(Cauchy inequalities)
\[\|F_{\theta_i}\|_{D(r-\sigma, s)}
\le c\sigma^{-1}\|F\|_{D(r,s)},\]
and
\[ \|F_{I}\|_{D(r,\frac 12s)}\le
2\frac 1{s^2} \|F\|_{D(r, s)},\ \ \
\|F_{z_n}\|_{D(r,\frac 12s)}\le
2\frac {n^ae^{n\rho}}{s} \|F\|_{D(r,s)}
\]
\end{Lemma}
Let $\{\cdot, \cdot\}$ is Poisson bracket of smooth functions
\begin{equation}
\{F, G\}=
\sum (\frac {\partial F}{\partial \theta_i}
\frac {\partial G}{\partial I_i}-
\frac {\partial F}{\partial I_i}
\frac {\partial G}{\partial \theta_i})+\sum_{i\in {\Bbb N}}
\langle \frac{\partial F}{\partial z_i}, {\rm i} J_{d_i}\frac{\partial G}{\partial z_i}
\rangle ,
\end{equation}
where $J_{d_i}$ are standard symplectic matrix in ${\Bbb R}^{d_i}$.
\begin{Lemma}\label{FG}
If
\[ \|X_F\|_{r,s}<\epsilon',
\|X_G\|_{r,s}<\epsilon'',\]
then
\[\|X_{\{F, G\}}\|_{ r-\sigma,\eta s}
\;<\;{\rm c}\; \sigma^{-1} \eta^{-2}\epsilon'\epsilon'', \quad \eta\ll 1.\]
\end{Lemma}
\Proof
Note that
\begin{eqnarray} \frac d{dz_n}\{F,G\}& &=\langle F_{\theta z_n}, G_I\rangle
+\langle F_\theta,
G_{Iz_n}\rangle
-\langle F_{Iz_n}, G_\theta\rangle -\langle F_I, G_{\theta z_n}\rangle \nonumber\\
& & +\sum_{i\in {\Bbb N}}(\langle F_{z_iz_n}, J_{d_i}G_{z_i}\rangle +
\langle F_{z_i}, J_{d_i}G_{z_iz_n}\rangle )
\end{eqnarray}
Since
\begin{eqnarray}
\|\langle F_{\theta z_n}, G_I\rangle \|_{D(r-\sigma,s)}& \;<\;{\rm c}\; &
\sigma^{-1} \|F_{z_n}|\|\cdot \|G_y\|
\nonumber\\
\|\langle F_\theta, G_{Iz_n}\rangle \|_{D(r-\sigma, \frac 12 s)}& \;<\;{\rm c}\; &
s^{-2} \|F_{\theta}\|\cdot \|G_{z_n}\|
\nonumber\\
\|\langle F_{Iz_n}, G_\theta\rangle \|_{D(r, \frac 12 s)}& \;<\;{\rm c}\; &
s^{-2} \|F_{z_n}\|\cdot \|G_\theta\|
\nonumber\\
\|\langle F_I, G_{\theta z_n}\rangle \|_{D(r-\sigma, s)}& \;<\;{\rm c}\; &
\sigma^{-1} \|F_{I}\|\cdot \|G_{z_n}\|
\nonumber\\
\|\langle F_{z_iz_n}, J_{d_i}G_{z_i}\rangle \|_{D(r, \frac 12s)}& \;<\;{\rm c}\; &
s^{-1} \|F_{z_n}\|\cdot \|G_{z_i}\| i^ae^{i\rho}
\nonumber\\
\|\langle F_{z_iz_n}, J_{d_i}G_{z_i}\rangle \|_{D(r, \frac 12s)}& \;<\;{\rm c}\; &
s^{-1} \|F_{z_n}\|\cdot \|G_{z_i}\|i^ae^{i\rho}
\end{eqnarray}
it follows from the definition of the weighted norm(see (\ref{weightednorm})),
that
\[\|X_{\{F, G\}}\|_{ r-\sigma, \eta s}
\;<\;{\rm c}\; \sigma^{-1}\eta^{-2}\epsilon'\epsilon''.\]
In particular, if $\eta\sim \epsilon^{\frac 13}, \epsilon', \epsilon''\sim
\epsilon$, we have $\|X_{\{F, G\}}\|_{ r-\sigma, \eta s}\sim
\epsilon^{\frac 43}$.
\nolinebreak\hfill\rule{2mm}{2mm
\begin{Lemma}\label{extension1}
Let $\Cal O$ be a compact set in ${\Bbb R}^d$ for which (\ref{ODC}) holds.
Suppose that
$f(\xi)$ and $\omega(\xi)$
are $C^m$ Whitney-smooth function in $\xi\in \Cal O$ with $C_W^m$ norm
bounded by $L$.
Then
\[g( \xi)\equiv\frac{f(\xi)}{\langle k, \omega(\xi)\rangle }\]
is $C^m$ Whitney-smooth in $\Cal O$
with\footnote{Recall the definition in (\ref{Falpha}).}
\[\|g\|_{\Cal O} \;<\;{\rm c}\; \gamma^{-{\rm c}}|k|^{\rm c} L,\]
\end{Lemma}
\Proof
The proof follows directly from the definition of the Whitney's
differentiability.
\nolinebreak\hfill\rule{2mm}{2mm
A Similar lemma for matrices holds:
\begin{Lemma}\label{extension2}
Let $\Cal O$ be a compact set in ${\Bbb R}^d$ for which (\ref{ODC}) holds.
Suppose that
$B(\xi), A_i(\xi)$
are $C^m$ Whitney-smooth matrices and $\omega(\xi)$ is
a Withney-smooth function in $\xi\in \Cal O$ bounded by $L$.
Then
\[C(\xi)=B M^{-1},\]
is $C^m$ Whitney-smooth
with
\[\|F\|_{\Cal O} \;<\;{\rm c}\; \gamma^{-{\rm c}}|k|^{\rm c} L,\]
where $M$ stands for either
$\langle k, \omega\rangle I_{d_i}+ A_iJ_{d_i}$ if
$B$ is $(d_i\times d_i)$-matrix, or
$\langle k, \omega\rangle I_{d_id_j}+(A_iJ_{d_i})
\otimes I_{d_j}-I_{d_i}\otimes (J_{d_j}A_j)$
if $B$ is $(d_id_j\times d_id_j)$-matrix,
\end{Lemma}
For a $N\times N$ matrix $ M=(a_{ij})$,
we denote by $| M|$ its
determinant. Consider $M$ as a linear operator on
$(R^N, |\cdot|)$ where $| x|=\sum |x_i|$. Let
$\|M\|$ be its operator norm.
It is known $\| M \|$ is equivalent to norm $ \max |a_{ij}|$.
Since a constant depends only on the space dimension and two fixed
norms is irrelevant,
we will simply denote $\|M\|= \max |a_{ij}|$.
\begin{Lemma}\label{3}
Let $M$ be a $N\times N$ non-singular matrix with
$\|M\| \;<\;{\rm c}\; |k|$, then
\[\{\omega: \|M^{-1}\|>h\}
\subset\{\omega: |\det M| \;<\;{\rm c}\; \frac{|k|^{N-1}}h\}
\]\end{Lemma}
\Proof
Firstly, we note that if $M$ is a nonsingular $N\times N$ matrix with
elements bounded by $|m_{ij}|\le m$, by Cramer rule, the inverse of
$M$ is $M^{-1}= \frac 1{|M|}{\rm adj}M$. Thus
\[\|M^{-1}\| \;<\;{\rm c}\; \frac {m^{N-1}}{|{\rm det} M|} \]
where the constant depends on $N$.
In particular, if $m={\rm const} |k|, |{\rm Det} M|>
\frac{|k|^{N-1}}h $ then
\[ \|M^{-1}\| \;<\;{\rm c}\; h.\]
This proofs the lemma.
\nolinebreak\hfill\rule{2mm}{2mm
In order to estimate the measure of $\Cal R^{\nu+1}$, we need the
following lemma, which has been proven in \cite{XYQ} \cite{Y2}.
A similar estimate is also used by Bourgain \cite{B2}.
\begin{Lemma}\label{A.1}
Suppose that $g(u)$ is a $C^m$ function on the closure
$\bar I$, where \ $I\subset R^1$\ is a finite interval.
Let\ $I_h=\{u |\ \ \ |g(u)|<h \},\ \ \ h>0 .$\ If for
some constant $d >0$,
$|g^{(m)}(u)| \ge d $ for all $u \in I$, then
$ {\rm meas}\;(I_h) \le ch^{\frac1{m}}$
where
$ c=2(2+3+\cdots +m+d^{-1}).$
\end{Lemma}
For the proof of Lemma \ref{regularity}, we need the following
\begin{Lemma}\label{sum1}
\[ \sum_{j\in {\Bbb Z}}e^{-|n-j|r+\rho |j|}\le Ce^{\rho |n|},
\sum_{j, n\in {\Bbb Z}}|q_j|e^{-|n-j|r+ |n|\rho}\le C|q|_\rho\]
if $\rho <r, q\in \Cal Z_{\rho}$ where $C$ depends on $r-\rho$.
\end{Lemma}
\begin{Lemma}\label{sum2}
\[ \sum_{j\in {\Bbb Z}}(1+|n-j|)^{-K} |j|^a \;<\;{\rm c}\; |n|^a,
\sum_{j, n\in {\Bbb Z}}|q_j|(1+|n-j|)^{-k} |n|^a\le C|q|_a\]
if $K>a+1, q\in \Cal Z_{a,\rho=0}$ where $C$ depends on $K-a-1$.
\end{Lemma}
\medskip\noindent
The proofs of the above two lemmata are elementary and we omit them.
\bigskip\noindent
\noindent
{\it Proof of Lemma \ref{regularity}}:
Here we give a direct proof.
It is clearly enough to consider the case of $f(u)$ being a monomial $u^{N+1}$ for
some $N\ge 1$.
From (\ref{tildeG}), one can see that the regularity of $G$ implies
the regularity of $\tilde G$. In the following, we shall give the proof
for $G$.
\smallskip\noindent
Suppose that the potential $V(x)$ is analytic
in $|{\rm Im} x|<r$ (respectively, belongs to
Sobolev space $H^K$)
then the eigenfunctions are analytic
in $|{\rm Im} x|<r$ (respectively, belong to $H^{K+2}$). If we let
$\phi_i(x)=\sum a_i^ne^{{\rm i}\langle n, x\rangle}\; $ then(see, e.g.,\cite{CW})
\[|a_{i}^{n}| \;<\;{\rm c}\; e^{-|i-n|r} \quad {\rm respectively}
\quad |a_{i}^{n}| \;<\;{\rm c}\; (1+|n-i|
^{-K-2}).\]
Recall that
\[G(q)=\sum_{i_0,\cdots ,i_N}C_{i_0\cdots i_N}\frac{q_{i_0}\cdots q_{i_N}}
{\sqrt{\lambda_{i_0}\cdots \lambda_{i_N}}}\]
where
\[C_{i_0\cdots i_N}=
\int_{T^1}\phi_{i_0}\cdots\phi_{i_N}dx=\sum_{n_0+n_1+\cdots+n_N=0}
(\prod_{s=0}^Na_{i_s}^{n_s}),
\]
with
$|a_{i_s}^{n_s}| \;<\;{\rm c}\; e^{-|i_s-n_s|r}$ (respectively,
$|a_{i_s}^{n_s}| \;<\;{\rm c}\; (1+|n_s-i_s|
^{-K-2})$.
In what follows, we assume either $a=0, \rho>0$ or $a>0, \rho=0$.
Since
\[G_{q_j}=(N+1)\sum_{i_1,\cdots, i_N }
C_{ji_1\cdots i_N}
\frac{q_{i_1}\cdots q_{i_N}}
{\sqrt{\lambda_{j}\lambda_{i_1}\cdots \lambda_{i_N}}}
\]
it follows that
\begin{eqnarray}
& &\|G_q\|_{a+\frac12, \rho}= \|G_{q_0}\|+\sum_{j\ge 1}
|G_{q_j}| |j|^{a+\frac 12}e^{j\rho}\nonumber\\
& \;<\;{\rm c}\; &\sum_{j,i_1,\cdots,i_N,\atop
n_0+\cdots +n_N=0}|a_{j}^{n_0}|j^{a}e^{|j|\rho}
(\prod_{s=1}^N|a_{i_s}^{n_s}q_{i_s}|)
\nonumber\\
& \;<\;{\rm c}\; &
\sum_{j,i_1, \cdots, i_N;\atop n_0+\cdots +n_N=0}(1+|j-n_0|)^{-N}|j|^a
e^{|j|\rho-|n_0-j|r}
\left(\prod_{s=1}^N(1+|n_s-i_s|)^{-K-2}
e^{-|n_s-i_s|r}|q_{i_s}|\right)\nonumber\\
& \;<\;{\rm c}\; &
\sum_{i_1, \cdots, i_N;\atop n_0+\cdots +n_N=0}|n_0|^ae^{|n_0|\rho}
\left(\prod_{s=1}^N(1+|n_s-i_s|)^{-K-2}e^{-|n_s-i_s|r}|q_{i_s}|\right)\nonumber\\
& \;<\;{\rm c}\; &
\sum_{i_1, \cdots, i_N; \atop n_1, \cdots,n_N}(|\sum_{s=1}^Nn_s|)^a
e^{|\sum_{s=1}^N n_s|\rho}
\left(\prod_{s=1}^N(1+|n_s-i_s|)^{-K-2}e^{-|n_s-i_s|r}|q_{i_s}|\right)\nonumber\\
& \;<\;{\rm c}\; &
\sum_{i_1, \cdots, i_N;\atop
n_1,\cdots,n_N}
\left(\prod_{s=1}^N(1+|n_s-i_s|)^{-K-2} |n_s|^ae^{-|n_s-i_s|r+|n_s|\rho}|q_{i_s}|\right)\nonumber\\
& \;<\;{\rm c}\; &
\sum_{i_1, \cdots, i_N}
\left(\prod_{s=1}^N|i_s|^a e^{|i_s|\rho}|q_{i_s}|\right)\nonumber\\
& \;<\;{\rm c}\; &
\prod_{s=1}^N
\left(\sum_{i_s}|i_s|^ae^{|i_s|\rho}|q_{i_s}|\right) \;<\;{\rm c}\; |q|_{a,\rho}^N.
\end{eqnarray}
\nolinebreak\hfill\rule{2mm}{2mm
|
\section{Introduction}
Let ${\cal B} = \{B_t\}_{t\in G}$ be a Fell bundle over a locally compact
group $G$ (see \cite{FD} for a comprehensive treatment of the theory
of Fell bundles, also referred to as $C^*$-algebraic bundles). We
denote by $L_2({\cal B})$ the right Hilbert $B_e$--module obtained by
completing the space $C_c({\cal B})$ of all continuous compactly supported
sections of ${\cal B}$, under the $B_e$--valued inner product
({\it cf.}~ \scite{E}{Section 2}, \scite{N}{2.2})
given by
$$
\<\xi,\eta\>_{B_e} =
\int \xi(t)^*\eta(t) \,d t \for \xi,\eta\in C_c({\cal B}).
$$
We should warn the reader that our notation for $L_2({\cal B})$ differs
from that used in \cite{N}.
The \stress{left-regular representation} of ${\cal B}$ is the map ({\it cf.}~
\scite{E}{2.2})
$$
\Lambda : {\cal B} \to {\cal L}(L_2({\cal B})),
$$
where ${\cal L}(L_2({\cal B}))$ indicates the $C^*$-algebra of all adjointable
operators \scite{JT}{1.1.7} on $L_2({\cal B})$, given for any $t$ in $G$
and any $b_t$ in $B_t$, by
$$
\Lambda(b_t) \xi \calcat s = b_t \xi(t^{-1} s)
\for \xi\in L_2({\cal B}), \quad s\in G.
$$
Let $C^*({\cal B})$ be the cross-sectional $C^*$-algebra of ${\cal B}$ ({\it cf.}~
\scite{FD}{VIII.17.2}) defined to be the enveloping $C^*$-algebra of
the Banach *-algebra $L_1({\cal B})$ formed by the integrable sections
\scite{FD}{VIII.5.2}.
The integrated form of $\Lambda$, which we also denote by $\Lambda$,
is the *-homomorphism
$$
\Lambda : C^*({\cal B}) \to {\cal L}(L_2({\cal B}))
$$
specified by setting $\Lambda(f) \xi = f * \xi$ ({\it cf.}~
\scite{N}{2.10}) for all $f$ in the dense subalgebra
$C_c({\cal B})\subseteq C^*({\cal B})$, and all $\xi\in C_c({\cal B})\subseteq L_2({\cal B})$.
Suppose that we are given a *-representation ({\it cf.}~
\scite{FD}{VIII.8.2 and 9.1}) $\pi$ of ${\cal B}$ on a Hilbert space ${\cal H}$,
i.e, a map $\pi:{\cal B}\to{\cal B}({\cal H})$ that is linear on each fiber, that
satisfies
\zitemno = 0\iItem {\zitemcntr} $\pi(b)\pi(c) = \pi(bc)$,
\Item {\zitemcntr} $\pi(b)^* = \pi(b^*)$,
\medskip\noindent for each $b,c\in{\cal B}$, and that is
\stress{continuous} in the sense that for each $u\in{\cal H}$, the map
$$
b\in{\cal B} \mapsto \pi(b)u\in{\cal H}
$$
is continuous in the norm of ${\cal H}$.
We may then form the representation $\pi_\lambda$ of ${\cal B}$ on $L_2(G)\*{\cal H} =
L_2(G,\H)$ by setting
$$
\pi_\lambda(b_t) = \lambda_t\*\pi(b_t) \for t\in G\for b_t\in B_t,
$$
where $\lambda_t$ refers to the left-regular representation of $G$
on $L_2(G)$.
We will also denote by
$$
\pi_\lambda : C^*({\cal B}) \to {\cal B}(L_2(G,\H))
$$
its integrated form \scite{FD}{VIII.11.2, 11.4, 17.2}.
Generalizing \scite{E}{2.3}, Ng defines in \scite{N}{2.11}
the \stress{reduced cross-sectional $C^*$-algebra} of ${\cal B}$, denoted
$C^*_r({\cal B})$, to be $\Lambda(C^*({\cal B}))$. Ng also proposes an
alternative notion of reduced algebra, namely
$$
C^*_R({\cal B}) := \pi_\lambda(C^*({\cal B})),
$$
where $\pi$ is any faithful *-representation of ${\cal B}$ on a Hilbert
space ${\cal H}$.
There exists (see below) a unique surjective *-homomorphism
$\Psi: C^*_R({\cal B}) \to C^*_r({\cal B})$ such that the diagram
$$
\comutriang{C^*_{R}({\cal B})}
$$
commutes. Ng thus introduced the notion of \stress{proper} Fell
bundles ({\it cf.}~ \scite{N}{2.15}) to single out those for which $\Psi$ is
injective. It is noticed in \cite{N} that Theorem 3.3 in
\cite{E} implies that Fell bundles over discrete groups are
automatically proper. It is also shown that saturated Fell bundles
are always proper \scite{N}{2.17}, as well as those whose underlying
group is compact \scite{N}{A.3}.
It is the purpose of this note to show that all Fell bundles are
proper and hence that the alternative reduced algebra $C^*_R({\cal B})$
proposed by Ng always coincides with $C^*_r({\cal B})$.
One of the main consequences is that the properness hypothesis
required in the main result of \cite{N} (Proposition 3.9) becomes
superfluous and hence we conclude that all Fell bundles satisfying the
\stress{approximation property} (Definition 3.6 in \cite{N}; see
also \scite{E}{4.4}) are amenable in the sense that $\Lambda$ is
an isomorphism from $C^*({\cal B})$ to $C^*_r({\cal B})$.
\section{Preliminaries}
Let us fix, throughout, a *-representation $\pi:{\cal B}\to{\cal B}({\cal H})$.
Restricting $\pi$ to $B_e$ we may view ${\cal H}$ as a left $B_e$--module
and hence we may form the tensor product
$L_2({\cal B})\*_{B_e}{\cal H}$
({\it cf.}~ \scite{R}{5.1}, \scite{JT}{1.2.3}), which is a
Hilbert space under the inner product defined by
$$
\<\xi\*u,\eta\*v\> =
\<u,\pi\(\<\xi,\eta\>_{B_e}\) v\>
\for \xi,\eta\in L_2({\cal B})\for u,v\in{\cal H}.
$$
\state Proposition
\label \PropoV
{\rm(Lemma 2.4 in \cite{N})}.
There exists an isometry
$$
V: L_2({\cal B})\*_{B_e}{\cal H} \to L_2(G,\H),
$$
such that for all $\xi\in L_2({\cal B})$, $u\in{\cal H}$, and $t\in G$ one has
$$
V(\xi\*u)\calcat t = \pi(\xi(t))u.
$$
\medbreak \noindent {\it Proof.\enspace }
It is obvious that $V$ is balanced with respect to the corresponding
actions of $B_e$ and hence it is well defined on the algebraic tensor
product
$L_2({\cal B})\odot_{B_e}{\cal H}$. Now let
$\xi,\eta\in L_2({\cal B})$, and $u,v\in{\cal H}$. We have
$$
\<V(\xi\*u),V(\eta\*v)\> =
\int \<\pi(\xi(t))u,\pi(\eta(t))v\> \,d t =
\<u,\pi\(\int\xi(t)^*\eta(t)\,d t\)v\> \$=
\<u,\pi(\<\xi,\eta\>_{B_e})v\> =
\<\xi\*u,\eta\*v\>,
$$
from which all of the remaining details follow.
\proofend
It should be observed that $V$ is not necessarily surjective. In
fact, note that the vector $V(\xi\*u)\calcat t$, mentioned above, lies
in $\pi(B_t){\cal H}$ which is often a proper subset of ${\cal H}$. This is
related to the notion of \stress{saturated} representations
\scite{N}{Definition 2.5} and is one of the main stumbling blocks we
must overcome in order to achieve our goals.
\state Proposition
\label \TensorId
{\rm(Lemma 1.3 in \cite{N})}.
If $\pi|_{B_e}$ is injective then so is
the *-homomorphism
$$
T \in {\cal L}(L_2({\cal B})) \longmapsto T\*1 \in {\cal B}(L_2({\cal B})\*_{B_e}{\cal H}).
$$
\medbreak \noindent {\it Proof.\enspace } Suppose that $T\*1=0$. Then, for all
$\xi,\eta\in L_2({\cal B})$, and $u,v\in{\cal H}$ we have
$$
0 =
\<(T\*1)(\xi\*u),\eta\*v\> =
\<u,\pi\(\<T(\xi),\eta\>_{B_e}\) v\>.
$$
Since $u$ and $v$ are arbitrary, and $\pi$ is supposed injective on
$B_e$, this implies that $\<T(\xi),\eta\>_{B_e}=0$ for all $\xi$ and
$\eta$, which in turn gives $T=0$.
\proofend
In particular, when $\pi|_{B_e}$ is injective, we have by
\lcite{\TensorId} that $C^*_r({\cal B})$ is isomorphic to the algebra
$\Lambda(C^*({\cal B}))\*1$ of operators on the Hilbert space
$L_2({\cal B})\*_{B_e}{\cal H}$.
\state Proposition
\label \SquareDiagram
For any $b\in {\cal B}$
the diagram
$$
\matrix{
&{\scriptstyle \Lambda(b)\*1} \cr
L_2({\cal B})\*_{B_e}{\cal H} & {\hbox to 1cm{\rightarrowfill}} & L_2({\cal B})\*_{B_e}{\cal H} \cr\cr
{\scriptstyle V}\Big\downarrow&& \Big\downarrow {\scriptstyle
V}\cr\cr
L_2(G,\H) & {\hbox to 1cm{\rightarrowfill}} & L_2(G,\H) \cr
&{\scriptstyle \pi_\lambda(b)}\cr
}
$$
commutes.
\medbreak \noindent {\it Proof.\enspace }
Let $t\in G$ be such that $b\in B_t$. We then have for all $\xi\in
L_2(B)$, $u\in{\cal H}$, and $s\in G$ that
$$
V(\Lambda(b)\*1)(\xi\*u)\calcat s =
V\big(\Lambda(b)\xi\*u\big)\calcat s =
\pi\(\Lambda(b)\xi\calcat s\) u =
\pi(b\xi(t^{-1} s))u.
$$
On the other hand
$$
\pi_\lambda(b)V(\xi\*u)\calcat s =
\pi(b)\(V(\xi\*u)\calcat {t^{-1} s}\) =
\pi(b)\pi(\xi(t^{-1} s))u.
\proofend
$$
It follows that the same holds if, in place of the ``$b$'' in the
statement above, we substitute any $a\in C^*({\cal B})$, since the
corresponding representations at the level of $C^*({\cal B})$ are integrated
from those of ${\cal B}$.
\definition
({\it cf.}~ \cite{N}).
Given a *-representation $\pi: {\cal B} \to {\cal H}$ as above we shall denote
by $C^*_{R,\pi}({\cal B})$ the algebra $\pi_\lambda(C^*({\cal B}))$ of operators on
$L_2(G,\H)$.
When $\pi|_{B_e}$ is faithful, $C^*_{R,\pi}({\cal B})$ was proposed by Ng
\cite{N} as an alternative reduced cross-sectional $C^*$-algebra
for ${\cal B}$.
The first relationship between $C^*_{R,\pi}({\cal B})$ and $C^*_r({\cal B})$ is
given by:
\state Proposition
\label \Triangle
Suppose that $\pi|_{B_e}$ is injective. Then
for any $a\in C^*({\cal B})$ one has that
$
\[\Lambda(a)\] \leq
\[\pi_\lambda(a)\].
$
Therefore there exists a unique *-homomorphism
$\Psi: C^*_{R,\pi}({\cal B}) \to C^*_r({\cal B})$ such that the diagram
$$
\comutriang{C^*_{R,\pi}({\cal B})}
$$
commutes.
\medbreak \noindent {\it Proof.\enspace }
By \lcite{\SquareDiagram} we have that $\Lambda\*1$ is equivalent to
a subrepresentation of $\pi_\lambda$. Therefore
$$
\[\Lambda(a)\* 1\] \leq
\[\pi_\lambda(a)\].
$$
Now, by \lcite{\TensorId}, we have that
$ \[\Lambda(a)\* 1\] = \[\Lambda(a)\]$.
The existence of $\Psi$ now follows by routine arguments.
\proofend
\section{The main result}
As already indicated, we plan to prove that $\Psi$ is an isomorphism
under the hypothesis that $\pi|_{B_e}$ is injective. This is clearly
equivalent to proving that for any $a\in C^*({\cal B})$ one has that
$
\[\Lambda(a)\] =
\[\pi_\lambda(a)\].
$
The starting point is that, although $\Lambda\*1$ is but a
subrepresentation of $\pi_\lambda$, we may ``move it around'' filling out the
whole of the representation space for $\pi_\lambda$. What will do the
``moving around'' will be the \stress{right-regular representation} of
$G$, namely the unitary representation $\rho$ of $G$ on $L_2(G)$ given
by
$$
\rho_r(\xi)\calcat s = \Delta(r)^{1/2}\xi(sr),
$$
for $\xi\in L_2(G)$, and $r,s\in G$, where $\Delta$ is, as usual,
the modular function for $G$.
\state Proposition
\label \RhoCommute
For each $r\in G$,
\zitemno = 0\iItem {\zitemcntr} The unitary operator $\rho_r\*1$ on $L_2(G)\*{\cal H} = L_2(G,\H)$
lies in the commutant of $\pi_\lambda(C^*({\cal B}))$.
\Item {\zitemcntr} Consider the isometry
$$
V_r: L_2({\cal B})\*_{B_e}{\cal H} \to L_2(G,\H),
$$
given by $V_r = (\rho_r\*1) V$. Then for all $a\in C^*({\cal B})$ one
has $V_r (\Lambda(a)\*1) = \pi_\lambda(a) V_r.$
\Item {\zitemcntr} Let $K_r$ be the range of\/ $V_r$. Then $K_r$ is invariant
under $\pi_\lambda$ and the restriction of\/ $\pi_\lambda$ to $K_r$ is equivalent to
$\Lambda\*1$.
\medbreak \noindent {\it Proof.\enspace }
It is clear that $\rho_r\*1$ commutes with $\pi_\lambda(b_t) =
\lambda_t\*\pi(b_t)$ for any $b_t\in B_t$. It then follows that
$\rho_r\*1$ also commutes with the range of the integrated form of
$\pi_\lambda$, whence (i). The second point follows immediately from (i) and
\lcite{\SquareDiagram}. Finally, (iii) follows from (ii).
\proofend
Our next result is intended to show that the $K_r$'s do indeed fill
out the whole of $L_2(G,\H)$.
\state Proposition
\label \Density
Suppose that $\pi|_{B_e}$ is non-degenerate. Then
the linear span of\/ $\bigcup_{r\in G} K_r$ is dense in $L_2(G,\H)$.
\medbreak \noindent {\it Proof.\enspace } Let
$$
\Gamma = {\rm span}\{V_r(\xi\*u): r\in G,\ \xi\in C_c({\cal B}),\ u\in
{\cal H}\}.
$$
Since
$$
V_r(\xi\*u)\calcat t =
(\rho_r\*1)V(\xi\*u)\calcat t =
\Delta(r)^{1/2}V(\xi\*u)\calcat {tr} =
\Delta(r)^{1/2}\pi(\xi(tr))u
\for t\in G,
$$
and since we are taking $\xi$ in $C_c({\cal B})$ above, it is easy to see
that $\Gamma$ is a subset of $C_c(G,{\cal H})$. Our strategy will be to use
\scite{FD}{II.15.10} for which we must prove that:
\medskip
\item{(I)} If $f$ is a continuous complex function on $G$ and
$\eta\in\Gamma$, then the pointwise product $f\eta$ is in $\Gamma$;
\item{(II)} For each $t\in G$ the set $\{\eta(t):\eta\in \Gamma\}$
is dense in ${\cal H}$.
\medskip
The proof of (I) is elementary in view of the fact that $C_c({\cal B})$ is
closed under pointwise multiplication by continuous scalar-valued
functions \scite{FD}{II.13.14}.
In order to prove (II) let $v\in{\cal H}$ have the form
$v=\pi(b)u$, where $b\in B_e$ and $u\in{\cal H}$. By
\scite{FD}{II.13.19} let $\xi\in C_c({\cal B})$ be such that $\xi(e)=b$.
It follows that
$\eta_r:=V_r(\xi\*u)$ is in $\Gamma$ for all $r$.
Also note that ,
setting $r=t^{-1}$, we have
$$
\eta_{t^{-1}}(t) =
\Delta(t)^{-1/2}\pi(\xi(e))u =
\Delta(t)^{-1/2}\pi(b)u =
\Delta(t)^{-1/2}v.
$$
This shows that $v\in\{\eta(t):\eta\in \Gamma\}$. Since the set of
such $v$'s is dense in ${\cal H}$, because $\pi|_{B_e}$ is non-degenerate,
we have that (II) is proven.
As already indicated, it now follows from \scite{FD}{II.15.10} that
$\Gamma$ is dense in $L_2(G,\H)$. Since $\Gamma$ is contained in the
linear span of $\bigcup_{r\in G} K_r$, the conclusion follows.
\proofend
The following is our main technical result:
\state Lemma
\label \Mainlemma
For all $a\in C^*({\cal B})$ one has that $\[\pi_\lambda(a)\] \leq
\[\Lambda(a)\]$.
\medbreak \noindent {\it Proof.\enspace } We may clearly suppose, without loss of generality, that
$\pi$ is non-degenerate. By \scite{FD}{VIII.9.4} it follows that
$\pi|_{B_e}$ is non-degenerate as well.
Under this assumption we claim that for all $a\in C^*({\cal B})$ one has
that
$$
\Lambda(a) = 0 \quad =\!\!\Rightarrow \quad \pi_\lambda(a) = 0.
$$
In order to see this suppose that $\Lambda(a) = 0$.
Then for each $r\in G$ we have by \lcite{\RhoCommute}.(ii) that
$
\pi_\lambda(a) V_r = V_r (\Lambda(a)\*1) =0.
$
Therefore
$\pi_\lambda(a)=0$ in the range $K_r$ of $V_r$.
By \lcite{\Density} it folows that $\pi_\lambda(a)=0$, thus proving our
claim.
Define a map
$$
\varphi : C^*_r({\cal B}) \longrightarrow {\cal B}(L_2(G,\H))
$$
by
$\varphi (\Lambda(a)) := \pi_\lambda(a)$, for all $a$ in $C^*({\cal B})$. By the
claim above we have that $\varphi$ is well defined. Also, it is easy
to see that $\varphi$ is a *-homomorphism. It follows that
$\varphi$ is contractive and hence that for all $a$ in $C^*({\cal B})$
$$
\[\pi_\lambda(a)\] =
\[\varphi (\Lambda(a))\] \leq
\[\Lambda(a)\].
\proofend
$$
Our main results follow more or less immediately from
\lcite{\Mainlemma}:
\state Corollary
\label \AlwaysProper
Let ${\cal B}$ be any Fell bundle over a locally compact group $G$ and let
$\pi$ be a *-representation of ${\cal B}$ on the Hilbert space ${\cal H}$ such
that $\pi|_{B_e}$ is injective. Then the map
$\Psi: C^*_{R,\pi}({\cal B}) \to C^*_r({\cal B})$
defined above is an isomorphism. Therefore ${\cal B}$ is always proper in
the sense of Ng \cite{N}.
\state Corollary
Suppose that the Fell bundle ${\cal B}$ satisfies the approximation
property (Definition 3.6 in \cite{N}; see also \scite{E}{4.4}),
then ${\cal B}$ is amenable in the sense that $\Lambda$ is an isomorphism
from $C^*({\cal B})$ to $C^*_r({\cal B})$.
\medbreak \noindent {\it Proof.\enspace } Combine Proposition 3.9 in \cite{N} with
\lcite{\AlwaysProper}.
\proofend
The following generalizes \scite{P}{7.7.5} to the context of Fell
bundles:
\state Corollary
Let $\pi:{\cal B}\to{\cal B}({\cal H})$ be a representation of the Fell bundle ${\cal B}$
and let $\pi_\lambda$ be the representation of ${\cal B}$ on $L_2(G,\H)$ given by
$\pi_\lambda(b_t) = \lambda_t\*\pi(b_t)$, for $t\in G$, and $b_t\in B_t$.
Denote also by $\pi_\lambda$ the representation of $C^*({\cal B})$ obtained by
integrating $\pi_\lambda$. Then $\pi_\lambda$ factors through $C^*_r({\cal B})$.
Moreover, in case $\pi|_{B_e}$ is faithful, the representation of
$C^*_r({\cal B})$ arising from this factorization is also faithful.
\medbreak \noindent {\it Proof.\enspace }
Follows immediately from \lcite{\Triangle} and \lcite{\Mainlemma}.
\proofend
\references
\bibitem{E}
{R. Exel}
{Amenability for {F}ell Bundles}
{\sl J. Reine Angew. Math. \bf 492 \rm (1997), 41--73}
\bibitem{FD}
{J. M. G. Fell and R. S. Doran}
{Representations of *-algebras, locally compact groups, and Banach
*-algebraic bundles}
{Pure and Applied Mathematics, 125 and 126, Academic Press, 1988}
\bibitem{JT}
{K. Jensen and K. Thomsen}
{Elements of $K\!K$-Theory}
{Birkh\"auser, 1991}
\bibitem{N}
{C.-K. Ng}
{Reduced Cross-sectional $C^*$-algebras of $C^*$-algebraic bundles
and Coactions}
{preprint, Oxford University, 1996}
\bibitem{P}
{G. K. Pedersen}
{$C^*$-Algebras and their automorphism groups}
{Acad. Press, 1979}
\bibitem{R}
{M. A. Rieffel}
{Induced representations of $C^*$-algebras}
{\sl Adv. Math. \bf 13 \rm (1974), 176--257}
\endgroup
\bye
|
\section{PRELIMINARY ESTIMATES}
Preliminary numerical estimates based on the above formulae for the
two different cases of LHC~\cite{LHC} and HIDIF~\cite{HIDIF}
designs give the following encouraging results:
\begin{center}
{\bf LHC}
\end{center}
\begin{tabular}{lll}
Transverse Emittance, $\epsilon$ & = & $3.75$ mm mrad \\
Total Energy $E$ & = & $450$ GeV \\
T & = & $25$ nano sec.\\
b & = & $1.2$ mm \\
P & = & $3.39 \times 10^{-5}$ \\
\end{tabular}
\begin{center}
{\bf HIDIF}
\end{center}
\begin{tabular}{lll}
Transverse Emittance, $\epsilon$ & = & $13.5$ mm mrad \\
Kinetic Energy $E$ & = & $5$ GeV \\
T & = & $100$ nano sec. \\
b & = & $1.0$ mm \\
P & = & $2.37 \times 10^{-3}$ \\
\end{tabular}
\section{CONCLUSION}
These preliminary numerical results are encouraging because they
predict halo losses which seem under control. Indeed the HIDIF
scenario gives a total loss of beam power per meter which is about a
thousand higher than the LHC. However in both cases the estimated losses
appear much smaller than the $1$ Watt/m.
|
\section{Introduction}
As it is well known, the equation
\begin{eqnarray}
\frac{dp}{dx} + \lambda\; p=0\; ,
\label{a1}
\end{eqnarray}
whose general solution is
\begin{eqnarray}
p(x)=p_{o}\exp(-\lambda x)\; ,
\label{po}
\end{eqnarray}
has many applications. In general, Eq. (\ref{a1})
models systems related to an {\it extensive} context (in the thermodynamical sense).
Typical examples are the systems based on independent events,
for instance, the radioactive decay of noninteracting nucleus.
In this context, it is interesting to observe
that the solution of Eq. (\ref{a1}) has an entropic interpretation.
In fact, if the usual entropy,
\begin{eqnarray}
S = - \int_a^{b} p(x)\; \ln p(x) \;{\mbox d}x \; ,
\label{a2}
\end{eqnarray}
is maximized subject to the constraints
\begin{eqnarray}
\alpha= \int_a^{b} p(x)\; {\mbox d}x
\label{a3}
\end{eqnarray}
and
\begin{eqnarray}
\beta=\int_a^{b} x \;p(x)\; {\mbox d}x \; ,
\label{a4}
\end{eqnarray}
the solution of Eq. (\ref{a1}) for $a\leq \;x\;\leq\;b$ is obtained again.
The parameters $\alpha$ and $\beta$ are adjusted
in order to obtain $p_o$ and $\lambda$.
If $\alpha=1$, $p(x)$ can be interpreted as a probability.
However, this choice for $\alpha$ is not necessary because
all the conclusions presented here are independent of the $\alpha$ value.
Now, it is possible to ask how Eq. (\ref{a1})
must be generalized in order to describe
dependent events ({\it nonextensive} context).
As a guide to answer this question it will
be considered in this work a
generalization of the entropy (\ref{a2})
employed by Tsallis\cite{ts88} (see also Ref. \cite{curado})
in a nonextensive statistical mechanics.
The above choice is motivated by the fact that this
generalized entropy (Tsallis entropy) has been
applied successfuly in the discussion of many
situations where the nonextensivity plays an important role,
for instance, L\'evy superdiffusion\cite{zanette}
and correlated anomalous diffusion\cite{plastino,tsallis,borland},
turbulence in two-dimensional pure electron plasma\cite{boghosian},
dynamic linear response theory\cite{rajagopal1} and Green functions\cite{mendes},
perturbation and variation methods for calculation of thermodynamic
quantities\cite{lenzi},
low-dimensional dissipative systems\cite{lyra},
simulated annealing and optimization techniques\cite{moret} and
connection with quantum uncertainty\cite{rajagopal2}.
In other words, the main purpose of this work is to obtain generalizations of
Eq. (\ref{a1}) based on nonextensive Tsallis entropy,
giving physical applications.
More specifically, it is obtained a generalization of Eq. (\ref{a1})
by using Tsallis entropy (Section {\bf II}), and applications to
the motion of a particle in a fluid medium and chemical kinetics
are presented.
A further generalization for higher order nonlinear differential equations
is given (Section {\bf III}) and it is applied to discuss a
new connection between the
correlated anomalous diffusion equation
and the Tsallis entropy.
Moreover, it is introduced a WKB-like approximation in order to study
a family of nonlinear differential equations based on
Tsallis entropy (Section {\bf IV}), and finally this approximated method
is employed to the case of the Thomas-Fermi equation.
\section{First order nonlinear differential equation}
The Tsallis entropy can be written as
\begin{eqnarray}
S_q = - \int^{b}_{a} \frac{p(x) \left (
1- p(x)^{q-1} \right )}{1-q} \; {\mbox d}x \; .
\label{a5}
\end{eqnarray}
Furthermore, the constraint (\ref{a4}) is currently
substituted\cite{ts88,curado} by
\begin{eqnarray}
\beta = \int_{a}^{b} x\; p(x)^q \;{\mbox d} x
\label{a6}
\end{eqnarray}
and the constraint (\ref{a3}) remains unchanged.
When the entropy (\ref{a5}) is maximized subject to the
constraints (\ref{a3}) and (\ref{a6}), $p(x)$ can be written as
\begin{eqnarray}
p(x) = p_o \left[1-(1-q)\;p_o^{q-1}\;
\lambda x \right]^{1 / (1-q)} \; .
\label{a7}
\end{eqnarray}
This distribution generalizes the exponential one (see Eq. (\ref{po})).
In the above expressions the parameter $q \; \in \; {\cal R}$
characterizes the degree of nonextensivity, in particular
the entropy (\ref{a2}) and the constraint (\ref{a3})
are obtained as limiting case when $q \rightarrow 1$,
recovering the extensive case.
A remarkable fact is that Eq. (\ref{a7}) can be applied
directly to several contexts successfuly.
In addition to other works already cited, the study of
solar neutrinos\cite{kaniadakis} and Zipf law\cite{denisov}
are relevant examples of such applications.
By direct inspection it is verified that the above
generalization of the exponential satisfies the
nonlinear equation
\begin{eqnarray}
\frac{dp}{dx}+\lambda \; p^q=0
\label{a8}
\end{eqnarray}
subject to the initial condition $p(0) = p_o$.
This equation is the generalization of Eq. (\ref{a1})
based on the nonextensive Tsallis entropy.
When the notation $\lambda_{eff}= \lambda p^{q-1}$ is employed,
Eq. (\ref{a8}) can be interpreted as a decay with memory,
thus $q$ is in some sense related with the memory of the system.
\subsection{Applications}
Before generalizing Eq. (\ref{a8}) to higher orders
it is illustrative to present applications for it.
One of them describes the mean motion of a particle in a fluid medium
without external force. In this case, the motion equation can be written as
\begin{eqnarray}
m{\mbox d}v / {\mbox d}t = - b\, v^{q} ,
\label {delta1}
\end{eqnarray}
where $v$ and $b$
are respectively the velocity and the friction coefficient of the particle
relative to the medium and $m$ is the particle mass. Here, the parameter $q$
is related to the turbulent flow (nonextensive
behavior) and for $q=1$ (slow motion) the extensive behavior is recovered.
Another example comes from chemical kinetics.
In this case, the concentration $C_A$ of a given species $A$
obeys the empirical equation\cite {quimica}
\begin{eqnarray}
{{\mbox d}C_A}/{{\mbox d}t}=K \, C_{A}^{\alpha}\,
C_{B}^{\beta}\, C_{C}^{\gamma} \cdot\cdot\cdot ,
\label{delta2}
\end{eqnarray}
where $K$ is the reaction constant and
$\alpha$, $\beta$, $\gamma$, $\cdot \cdot \cdot$,
refer to the concentration of chemical species $A$, $B$, $C$, $\cdot \cdot \cdot$,
present in the reaction. In this expression, $\alpha$, $\beta$, $\gamma$,
$\cdot\cdot\cdot$, are respectively the order of the reaction with respect to
$A$, $B$, $C$, $\cdot\cdot\cdot$,
and the sum $\alpha+\beta+\gamma+ \cdot \cdot \cdot$ is the
overall order of the reaction.
In some cases the concentrations $C_B$, $C_C$, ... can be considered constant,
thus the above equation reduces to the form of Eq. (\ref{a8}) with $q=\alpha$.
In this way, the parameter $q$ becomes the order of the reaction for the species $A$.
\section{Family of nonlinear differential equations}
The natural generalization of Eq. (\ref{a1}) for higher order differential
equations, in the sense that $p=p_{o}\exp(-\lambda x)$ is a particular solution,
is the family of linear differential equations of
arbitrary order with constant coefficients.
The $N$-order element of this family is
\begin{eqnarray}
\sum_{n=0}^{N}a_n \frac{{\mbox d}^n p}{{\mbox d} x^n} = 0
\label{ta}
\end{eqnarray}
and the corresponding algebraic equation for $\lambda$ is
\begin{eqnarray}
\sum_{n=0}^N a_n(-\lambda)^n = 0 \;\; .
\label{tb}
\end{eqnarray}
In view of the previous remarks, the generalization of these ideas
in the Tsallis context
is based on the replacement of the particular solution $p(x)=p_{o}\exp(-\lambda x)$ by
$p(x) = p_o[1 - (1-q) p_o^{q-1}\lambda x]^{1/(1-q)}$.
Thus,
\begin{eqnarray}
\sum_{n=0}^{N}a_n\frac{\mbox d^n}{ {\mbox d}x^n}p^{(N-n)(q-1)+1} = 0
\label{tc}
\end{eqnarray}
represents the nonlinear ordinary
differential equation of constant coefficients that generalizes
Eq. (\ref{ta}) and the corresponding generalization of Eq.(\ref{tb}) is
\begin{eqnarray}
a_N(-\lambda)^{N} \prod_{j=0}^{N-2}[(j+1)q-j] +
\sum_{n=1}^{N-1} a_n(-\lambda)^n \prod_{j=N-n-1}^{N-2}[(j+1)q-j] + a_0 = 0
\;\;.
\label{td}
\end{eqnarray}
This equation is applicable for $N\geq 2$ and when $N=1$
Eq.(\ref{td}) must be replaced by $a_1(-\lambda) + a_0=0$.
Without loss of generality the $N$-th term of Eq. (\ref{tc}),
$\mbox d^N p/\mbox dx^N$,
was considered linear in $p(x)$.
As expected, Eqs. (\ref{tc}) and (\ref{td}) reduce respectively to
Eqs. (\ref{ta}) and (\ref{tb}) in the limit $q\rightarrow 1$. Furthermore,
it is important to remark that a superposition of particular solutions
of Eq. (\ref{tc}) is not another solution of this equation, in contrast with the
case of Eq. (\ref{ta}).
For instance, the generalization of circular and hyperbolic functions
based on the superposition of Eq. (\ref{a7}) \cite{borges} is not a solution
of Eq. (\ref{tc}).
Thus, in the following applications only particular solutions
previously described will be considered.
\subsection{Application to correlated anomalous diffusion}
It was recently shown that the one-dimensional correlated anomalous diffusion
equation (generalized nonlinear Fokker-Planck equation without external force),
\begin{eqnarray}
\frac{\partial \phi}{\partial t }=\frac{\partial^2\phi^{\nu}}{\partial x^2} \;\;\;
(\nu\in \cal R)\; ,
\label{te}
\end{eqnarray}
has a solution related to the Tsallis entropy \cite{plastino,tsallis}.
This solution is a generalization of the Gaussian one, reducing to it
in the limit $q\rightarrow 1$.
On the other hand, by using the generalization developed above
for linear ordinary differential equations,
a new relation between Tsallis entropy
and correlated anomalous diffusion equation can be obtained.
The new connection is based on the substitution of
\begin{eqnarray}
\phi(x,t) = T(t)\, X(x)
\label{delta4}
\end{eqnarray}
into Eq. (\ref{te}).
This separation of variables leads to
${ {\mbox d} T}/{ {\mbox d} t} = -\sigma \;T^{\nu}$
and
${ {\mbox d}^2 Y}/{ {\mbox d}x^2} = -\sigma \;Y^{1/\nu}$,
where $\sigma$ is a separation variable constant and $Y = X^{\nu}$.
Since these equations are members of the family
ruled by Eq. (\ref{tc}), a new connection
between the one-dimensional correlated anomalous diffusion equation and the
Tsallis entropy is established.
Furthermore, as a consequence of this general procedure,
a set of particular solutions of the
anomalous correlated diffusion equations is obtained,
namely,
\begin{eqnarray}
\phi (x,t)=T_{0} \left[ 1-(1-\nu)\;T_{0}^{\nu-1}\; \sigma \;t \right]^{1/(1-\nu )}\; X_{0}
\left[ 1+\left(\frac{1-\nu}{2\nu}\right) X_{0}^{(1- \nu )/2} \lambda\; x \right]^{2/(\nu -1)} \;\;,
\label{delta5}
\end{eqnarray}
where $T_0$ and $X_{0}$ are constants and $\lambda$ is the solution of the equation
$(1+\nu) \lambda^{2}+2 \nu \sigma=0$.
Note that the above procedure leads to complex solutions for $\sigma >0$
and it is a natural extension of the method of separation of variables
applied to the usual diffusion equation ($\nu =1$).
The previous method can be easily applied to
the generalized diffusion equation recently proposed
by Tsallis and Bukman\cite{tsallis},
\begin{eqnarray}
\frac {\partial {\phi}^{\mu}} {\partial t}=\frac {\partial ^2 {\phi}^{\nu}}{\partial x^2}
\;\;\;\; (\mu, \nu \in {\cal{R}})\; .
\label{delta7}
\end{eqnarray}
In fact, it is sufficient to replace $\phi ^{\mu}$ by $\psi$ and $\nu$ by $\nu / \mu$.
In general, the connection between the porous media diffusion
equation and Tsallis entropy, based on separation of variables and the family
of nonlinear partial differential equation (Eq. (\ref{tc})),
can be easily extended for other nonlinear
partial differential equations.
\section{ Nonlinear differential equations and WKB-like approximation}
Eq. (\ref{a8}) can be generalized further if we
allow $\lambda$ to become a function of $x$.
In this case, the solution of Eq. (\ref{a8}) with
$p(0)=p_o$ becomes
\begin{eqnarray}
\label{a9}
p(x)= p_o \left[ 1-(1-q)p_o^{q-1}
\int_0^x \lambda(z) {\mbox d}z \right ]^{1/(1-q)} \; .
\end{eqnarray}
In a similar way, we can allow that the constants
$a_n$ in Eq.(\ref{tc}) become functions of $x$.
This procedure leads to
\begin{eqnarray}
\sum_{n=0}^{N}a_n(x)\frac{ {\mbox d}^n}{ {\mbox d}x^n}
p^{(N-n)(q-1)+1}=0\; .
\label{ti}
\end{eqnarray}
As in the linear case ($q=1$) there are no general solutions for these equations.
Consequently, it is natural to perform approximated analyses to obtain some information
about the solutions of Eq.(\ref{ti}).
In the following discussions, among other possibilities,
a generalization of WKB method for $q\neq 1$
is developed explicitly for the $N=2$ case.
As it is well known, in the WKB method\cite{wkb}
approximate solutions of equation
${{\mbox d}^2p}/{{\mbox d}x^2}= f(x)p$
can be obtained when $f(x)$ is a slowly varying function.
In this case, it is employed an auxiliary function $g(x)$
defined by the relation
$p(x)=\exp{(g(x))}$.
>From the previous developments, a natural generalization
of these equations are respectively
\begin{eqnarray}
\frac{{\mbox d}^2p}{{\mbox d}x^2}= f(x)p^{2q-1}
\label{b11}
\end{eqnarray}
and
\begin{eqnarray}
p(x)=[1+(1-q)g(x)]^{1/(1-q)} \; .
\label{a13}
\end{eqnarray}
In terms of $g(x)$, Eq. (\ref{b11}) becomes
\begin{eqnarray}
[1+(1-q) g]\frac{{\mbox d}^2 g}{{\mbox d}x^2} + q
\left( \frac{{\mbox d} g}{{\mbox d}x}\right)^2 - f = 0 \; .
\label{a14}
\end{eqnarray}
Following again the usual WKB approach, the term with ${{\mbox d}^2 g}/{{\mbox d}x^2}$
is neglected in the first approximation, hence
\begin{eqnarray}
g(x)= \pm q^{-1/2} \int f(x)^{1/2} {\mbox d} x \; .
\label{a16}
\end{eqnarray}
Consistently, the validity condition of this approximation is
\begin{eqnarray}
\left| \frac {{\mbox d}^2 g}{{\mbox d}x^2}\right| \approx
\left| \frac {{{\mbox d} f}/{{\mbox d}x}}{2q^{1/2}f^{1/2}} \right|
\ll \left| \frac {f}{1+(1-q)g} \right| \;.
\label{a17}
\end{eqnarray}
Since Eq. (\ref{b11}) is nonlinear for $q\neq 1$, the
superposition of solutions is not a solution,
therefore the present development is indicated for situations where some
particular solutions can be considered as good approximations.
Notice also that the solutions (\ref{a16}) can be improved
through iterations (replacing successively improved Eq. (\ref{a16}) into Eq. ({\ref{a14})).
\subsection{Application to Thomas-Fermi equation}
To exemplify the previous development, we consider the
Thomas-Fermi equation for a free
atom\cite{landau}
\begin{eqnarray}
\frac{{\mbox d}^{2}y}{{\mbox d}x^2}= x^{-1/2} \; y^{3/2} \; .
\label{a19}
\end{eqnarray}
In the free neutral atom case the boundary conditions
are $y(0)=1$ and $y(\infty)=0$.
By comparing Eq. (\ref{a19}) with Eq. (\ref{b11}) it is
verified that $f(x)=x^{-1/2}$ and $q=5/4$.
Choosing the particular solution adjustable
to these conditions it becomes
\begin{eqnarray}
y(x)=\left(1+\frac{2}{3 \sqrt{5}}\; x^{3/4}\right)^{-4} \; .
\label{a20}
\end{eqnarray}
In this example, the validity condition (\ref{a17}) becomes
\begin{eqnarray}
1+\frac {3\sqrt{5}} {2}\; x^{-3/4} \ll 15 \; .
\label{a21}
\end{eqnarray}
This condition indicates that the approximation is better
for larger $x$.
Furthermore, Eq. (\ref{a20}) is in satisfactory
agreement with a numerical calculation
(see, for instance, Ref. \cite{landau}).
In a general context, when $f(x)$ is a smooth function, the
corresponding approximate solution becomes more accurate.
In particular, when $f(x)$ is a constant this solution becomes exact.
\section{Conclusions}
Summing up, the nonextensive concepts based on the Tsallis entropy
were employed to obtain a family of nonlinear
ordinary differential equations.
The first order equation of this family is a nonextensive
generalization of the exponential decay equation.
Moreover, by using a separation of variables procedure
and the above family of equations, a connection between
the correlated anomalous diffusion
equation and the Tsallis entropy is obtained.
In addition to this, for second order equations we presented a WKB-like
approach to obtain approximated solutions.
This procedure was used in the context of the
Thomas-Fermi equation for a free neutral atom, and it was shown that
the well known solution precisely correspond to $q=5/4$.
In general, the developments introduced in this work indicate that
many nonlinear effects are closely related with nonextensive
concepts in the Tsallis framework.
\acknowledgements
I. T. Pedron thanks CAPES (Brazilian Agency) for financial support.
|
\section{Modified WKB approximation for the hydrogen atom}
We start from the radial Schr\"odinger equation for the hydrogen atom
\begin{equation}
\left(-\frac{\hbar^{2}}{2m}\frac{d^{2}}{dr^{2}}-
\frac{e^2}{r}+V_{C}(r)\right)\Psi (r)=E \Psi (r)
\label{eq:Schrodrad}
\end{equation}
with $V_{C}(r)$ given by Eq.\ (\ref{eq:Zentrf1}).
Using the conventional WKB ansatz for the wave function
\begin{equation}
\Psi (r) =
\exp\left[\frac{i}{\hbar}\sum_{k=0}^{\infty}(-i\hbar)^k S_{k}(r)\right]
\label{eq:Ansatz}
\end{equation}
and expanding in powers of $\hbar$, we obtain for the quantities
\begin{equation}
y_{k}(r,E,L)=\partial S_{k}(r,E,L)/\partial r
\label{eq:defy}
\end{equation}
the recursive set of equations
\begin{eqnarray}
y_{0} & = & p(r,E,L) = \pm
\sqrt{2m(E-V_{\rm eff}(r))}\label{eq:dgl1} \\
y_{1}& = & -\frac{1}{2 y_{0}}\left(y'_{0}
+i\frac{L}{r^2}\right) \label{eq:dgl2}\\
y_{2m} & = &
-\frac{1}{2y_{0}}\left[y_{m}^2+y'_{2m-1}+2\sum_{k=1}^{2m-2}
y_{2m-k}y_{k}\right] \label{eq:dgl3} \\
y_{2m+1} & = &
-\frac{1}{2y_{0}}\left[y'_{2m}+2\sum_{k=1}^{2m-1}y_{2m+1-k}y_{k}\right]
\label{eq:dgl4}
\end{eqnarray}
where
\begin{equation}
V_{\rm eff}(r)= -\frac{e^2}{r}+\frac{L^{2}}{2m r^{2}}.
\label{eq:Zentrf2}
\end{equation}
and where $p(r,E,L)=y_{0}(r,E,L)$ is the classical momentum. Further, the
prime denotes differentiation with respect to $r$.
These equations yield two functions $y^{\left(\pm\right)}(r,E,L)$
depending on the choice of the sign of the momentum $p(r,E,L)$, and
the wave function is a linear combination of the form
\begin{equation}
\Psi\left(r,E,L\right) =
\sum\limits_{\sigma=\pm} c^{(\sigma)}{\rm exp} \left({\frac{i}{\hbar}
\int\limits_{r_{0}}^{r}dr\,
y^{(\sigma)}(r,E,L)}\right) \label{eq:Wfkt1}
\end{equation}
where
\begin{equation}
y(r,E,L)=\sum\limits_{k=0}^{\infty} \left(-i\hbar\right)^k y_{k}(r,E,L).
\end{equation}
The momentum $p(r,E,L)$ has a branch cut which is chosen conveniently
between the classical turning points
\begin{equation}
r_{1,2}=a\left(1\mp\epsilon\right),
\label{eq:umkp}
\end{equation}
where $a$ is the big axis and $\epsilon$ the eccentricity of the
ellipse in the Kepler problem. Dunham \cite{Dunham} has shown that
by choosing the initial point of
integration $r_{0}$ on the left side of the two classical turning points
and a contour avoiding
the turning points as indicated in Fig.\ 1a, the wave function becomes
\begin{eqnarray}
\Psi\left(r,E,L\right)=\left\{\begin{array}{l@{\ ,\quad}l}
c^{(-)}\left(\Psi^{(-)}+\Psi^{(+)}
\right) & r_{1}<r<r_{2} \\
c^{(-)}\Psi^{(-)} &
{\rm elsewhere}\label{eq:Wfkt2}
\end{array}\right.
\end{eqnarray}
with
\begin{equation}
\Psi^{(\pm)}(r,E,L)=\exp\left(\frac{i}{\hbar}\int_{r_{0}}^{r}dr\,
y^{(\pm)}(r,E,L)\right).\label{eq:Wfkt4}
\end{equation}
\vspace{-.5cm}
\begin{figure}
\begin{center}
\leavevmode
\epsfxsize=0.4 \textwidth
\epsfbox{pic4.eps}
\end{center}
\end{figure}
\begin{figure}[H]
\vspace{-1cm}
\begin{center}
\leavevmode
\epsfxsize=0.4 \textwidth
\epsfbox{pic3.eps}
\end{center}
\caption{ a) The complex $r$ plane with the classical turning points
$r_{1/2}$. Connecting points
of the classical allowed and forbidden regions one has to avoid the
turning points by integrating along the circles.
b) Deformation of the integration contour in the complex plane.}
\end{figure}
Since we search for a unique solution, we have to require that the
wave function is independent of whether one integrates above or below
the branch cut. This leads to the condition
\begin{equation}
\frac{i}{\hbar}\oint dr y\left(r,e,L\right)=2\pi i \left(n_{r}+1\right),
\end{equation}
where $n_{r}$ is a positive integer and the integration contour
encircles the branch cut. Using this equation one gets a
quantization of the energy which is related to the
Bohr-Sommerfeld rule.
To evaluate the contour integrals, we use a technique due to Sommerfeld
which exploits the fact that the $y_{k}(r,E,L)$ have only poles on the
positive real axis.
By this assumption we find
As indicated in Fig.\ 1b one has to
calculate integrals along the contours $C_{2}$ and $C_{3}$
instead of encircling the branch cut. To order $\hbar$ the integrals
are readily evaluated yielding
\begin{displaymath}
\frac{1}{2\pi\hbar}\oint dr\, \left(y_{0}+\frac{\hbar}{i}
y_{1}\right)
=-\frac{L}{\hbar}+\sqrt{-\frac{m
e^{4}}{2 E\hbar^2}}=n_{r}+1,
\end{displaymath}
which gives the exact energy
eigenvalues for the bound states of the hydrogen atom
\begin{equation}
E_{n}=-m e^4/2 \hbar^2 n^2
\label{eq:Rydberg}
\end{equation}
with the principal quantum number $n=n_{r}+l+1$. Corrections of higher
order in $\hbar$ coming from the contour integrals
over the functions $y_{k}$, $k\geq 2$ vanish exactly. To show this we
first investigate the analytical structure of $y_{0}$ and $y_{1}$ at
the origin. We find
\begin{equation}
y_{0}(r,E,L)=iLr^{-1}+O(r^0)
\end{equation}
while the power series expansion of $y_{1}$ begins with a linear
term. Consequently the expansion of $y'_{1}$ starts with a constant
term. Now, using
\begin{equation}
y_{2}=-\frac{y_{1}^2+y'_{1}}{2 y_{0}}
\end{equation}
one immediately sees that the expansion of $y_{2}$ begins with a
linear term and therefore the residue of $y_{2}$ at the origin is zero.
Since the recurrence relations (\ref{eq:dgl3}) and (\ref{eq:dgl4})
contain $y_{0}$ only in the denominator, it is easy to show by
induction that the Taylor series of all $y_{k}$ with $k \geq 2$ start
with linear or higher order terms. This implies that the integrals
along the contour
$C_{2}$ vanish for all $y_{k}$ with $k\geq 2$. In an analogous way
one can treat the integrals along the contour $C_{3}$ by replacing
$r$ by $1/u$ and remembering the additional factor $-1/u^{2}$
originating from the transformation of the integration measure. One
finds that the integrals along the contour $C_{3}$ also vanish for
all $y_{k}$ with $k\geq 2$.
Therefore the semiclassical energy quantization (\ref{eq:Rydberg}) is
exact to all orders in $\hbar$, while in the WKB approximation with Langer
modification higher order terms destroy the exactness of the energy
eigenvalues.
Next we consider the wave functions.
Disregarding quadratic and higher powers in $\hbar$ in
(\ref{eq:Wfkt2}) and (\ref{eq:Wfkt4}),
we arrive at an expression for the lowest order WKB wave functions
for $r$ on the positive real axis of the form
\begin{equation}
\Psi(r,E,L)=\frac{1}{2}{\rm
Re}\left[\Psi^{(-)}(r,E,L)+\Psi^{(+)}(r,E,L)\right]
\label{eq:mwwfkt1}
\end{equation}
with
\begin{equation}
\Psi^{(\pm)}(r,E,L)=\frac{c(E,L)}
{\sqrt{p(r,E,L)}}e^{\pm\left(\frac{i}{\hbar}\int_{r_{1}}^{r}dr\,
p -\frac{i}{2}\varphi-i\frac{\pi}{4}\right)},
\label{eq:mwwfkt2}
\end{equation}
where the additional phase $\varphi(r,E,L)$ arises from
the part of the centrifugal term in (\ref{eq:Zentrf1}) that is linear
in $\hbar$. In fact,
\begin{equation}
\varphi(r,E,L)= -\partial S_{0}(r,E,L)/\partial L
\end{equation}
is just the phase of the classical trajectory in the plane of motion of the
Kepler problem
in terms of which eq.\ (\ref{eq:dgl2}) can be written as
\begin{equation}
y_{1}=-\frac{y'_{0}}{2
y_{0}}-\frac{i}{2}\frac{\partial\varphi}{\partial r}.
\end{equation}
A representation of the WKB wave function as the real part of the
superposition of incoming and outgoing waves as in
Eq.\ (\ref{eq:mwwfkt1}) was introduced previously
by More and Warren
\cite{M&W} for the standard approach with LM. Since the undesirable
growing part of the wave function has a purely imaginary
coefficient, it is removed when the real part is taken. The normalization
$c(E,L)$ of the wave function is obtained from
\begin{equation}
\frac{1}{2}{\rm Re}\int_{r_{1}}^{r_{2}} dr\,
\Psi^{(+)}\Psi^{(-)}
=\frac{1}{4}\oint dr\, \Psi^{(+)}\Psi^{(-)}=1.\label{eq:norm2}
\end{equation}
which gives
\begin{equation}
c(E)^2 =\frac{2 m}{\pi\hbar^2}\frac{dE}{dn}.
\label{eq:norm3}
\end{equation}
More and Warren refer to the omission of the terms
$
\Psi^{(+)}\Psi^{(+)}+\Psi^{(-)}\Psi^{(-)}
$
in the normalization integral as
"restricted interference approximation".
Finally, we get for the WKB wave function of the hydrogen atom in the
classical accessible region between the two turning points
\begin{equation}
\Psi(r,E,L)=\frac{c(E)}{\sqrt{p}}\cos\left(\frac{1}{\hbar}
\int\limits_{r_{1}}^{r}dr\,
p -\frac{\varphi}{2}-\frac{\pi}{4}\right).
\end{equation}
We now compare the WKB wave functions with the exact ones. A typical
feature of WKB wave functions is the divergence at the classical turning
points. As can be seen from Fig.\ 2, this behavior is qualitatively
the same for the Langer modified expansion (LM), our systematic
$\hbar$-expansion WKB(SE), and poor man's WKB(PM) obtained when the
full centrifugal term (\ref{eq:Langerterm2}) is retained in the lowest
order equation. While the WKB(PM) wave function for the ground state
does indeed poorly, the main difference between the WKB(LM) and WKB(SE)
wave functions comes from the fact that the distance between the
turning points of the Langer modified wave functions is smaller. This
is just a consequence of the shift of the turning points due to the
LM. Therefore, between the turning points, our wave functions give a
better approximation to the exact ones.
For the s-states, the Langer modified wave functions are constructed
to vanish at $r=0$ and they have a divergence near the origin since the
left turning point is moved away from $r=0$ by the artificial $1/2$
added to the angular momentum number $l$.
Our wave function doesn't have the right
power law behavior near the origin
but there is only one divergence which is due to the right turning
point. Hence, we see that the
wave functions obtained from a systematic expansion in powers of
$\hbar$ without any {\it ad hoc} manipulation of the hydrogen problem
are at least as accurate as those obtained from the problem with
LM.
\begin{figure}
\begin{center}
\leavevmode
\epsfxsize=0.4 \textwidth
\epsfbox{plot1s.eps}
\end{center}
\caption{WKB wave functions and exact wave functions for the 1s ground
state. The WKB wave functions diverge at the turning points. $r$ is
measured in units of the Bohr radius $a_{0}$.}
\label{fig:fig3}
\end{figure}
\begin{figure}
\begin{center}
\leavevmode
\epsfxsize=0.4 \textwidth
\epsfbox{plot3s.eps}
\end{center}
\vspace{-1cm}
\end{figure}
\begin{figure}
\begin{center}
\leavevmode
\epsfxsize=0.4 \textwidth
\epsfbox{plot3p.eps}
\end{center}
\caption{WKB wave functions and exact wave functions for the excited
$3s$ and $3p$ states. Again the full, dashed, and dotted lines denote
the exact, WKB(LM), and the WKB(SE) wave functions, respectively.}
\label{fig:fig4}
\end{figure}
Finally, we calculate radial dipole matrix elements between states
with angular momentum $l$ and $l\pm 1$. Using the restricted
interference approximation, we have
\begin{eqnarray}
& R^{\pm}_{\Delta n}(E,l) = & \label{eq:matrelm} \\
& \frac{1}{4}\oint dr\, \Psi^{(+)}(r,E,L) r
\Psi^{(-)}(r,E+\Delta E,L \pm\hbar). & \nonumber
\end{eqnarray}
Expanding this in powers of $\hbar$, one finds for the leading order term
\begin{eqnarray}
R_{\Delta n}^{\pm\, (0)}(E,l) =
a_{0}\left(\frac{n^2}{\Delta
n^2}\frac{d}{d\epsilon}J_{\Delta n}(\Delta n\epsilon)\right. \\
\left. \pm
\frac{n}{\Delta n}\frac{\sqrt{1-\epsilon^2}}{\epsilon}J_{\Delta
n}(\Delta n\epsilon)\right)\nonumber
\end{eqnarray}
where $a_{0}$ is the Bohr radius,
$\epsilon=\left[1-(l/n)^2\right]^{1/2}$ the eccentricity, and
$J_{n}(z)$ a Bessel function.
Naccache \cite{Nac} has obtained this leading order term from the Heisenberg
correspondence principle.
The quantum correction of first order in $\hbar$ is found to read
\begin{eqnarray}
R_{\Delta n}^{\pm\, (1)}(E,l) = \frac{\Delta
n\omega(E)}{2}\frac{\partial}
{\partial E}R_{\Delta n}^{\pm\, (0)}(E,l)\label{eq:matrelm1Or}\\
+\frac{1\pm 1}{2}\frac{\partial}{\partial L}
R_{\Delta n}^{\pm \, (0)}(E,l)\nonumber
\end{eqnarray}
with the angular frequency $\omega(E)=\left[-8 E^3/ (m e^6)\right]^{1/2}$
of the Kepler problem. In Tab.\ 1 the semiclassical dipole elements
are compared with the exact ones for some spectral series.
We note that
for large $n$ and $l$ and small $\Delta n$ the WKB results give rather
accurate estimates of the exact values. This is expected from a semiclassical
approximation.
In summary, we have shown that a systematic semiclassical expansion of
the hydrogen problem about the Kepler problem yields remarkably
accurate results. In contrast to the common belief no modification of
the WKB expansion is necessary when the centrifugal potential term is
decomposed in the classical centrifugal potential and a quantum
correction. The same method can be employed to other problems with
radial symmetry.
The authors would like to thank Joachim Ankerhold and Phil Pechukas
for valuable discussions and acknowledge support by the SFB 276 of
the Deutsche Forschungsgemeinschaft (Bonn). Additional support was
provided by the Deutscher Akademischer Austausch\-dienst (DAAD).
\end{multicols}
\newpage
\widetext
\begin{table}
\begin{tabular}{|c|c|c|c|c|c|c|c|c|c|}
n & 2 & 3 & 4 & 5 & 6 & 7 & 8 & 9 & 10 \\ \hline\hline
1s-np & 1.090(1.290) & 0.512(0.517) & 0.339(0.306) & 0.257(0.209)
& 0.208(0.155) & 0.177(0.121) & 0.154(0.098) & 0.137(0.082)
& 0.124(0.069) \\
2p-nd & - & 4.542(4.748) & 1.816(1.71) & 1.104(0.975) & 0.802(0.662)
& 0.641(0.492) & 0.543(0.386) & 0.478(0.314) & 0.432(0.263) \\
4p-ns & - & - & - & 4.673(4.600) & 1.864(1.788) & 1.120(1.044)
& 0.794(0.718) & 0.614(0.539)
& 0.501(0.427)
\end{tabular}
\caption{WKB(SE) dipole matrix elements in units of Bohr's radius
$a_{0}$ and exact quantum mechanical values in parenthesis.}
\label{tab:tab1}
\end{table}
\begin{multicols}{2}
|
\section{Introduction}
The X-ray transient XTE~J2123--058 was discovered by the {\it RXTE}
satellite on 1998 June 27 (Levine, Swank and Smith 1998, \cite{L98}).
An optical counterpart was promptly identified by
Tomsick et al.\ (1998a, \cite{T98a}). The discovery of apparent
Type-I X-ray bursts (Takeshima and Strohmayer 1998, \cite{TS98})
indicated that the compact object was a neutron star. Interest in the
object increased dramatically when Casares et al.\ (1998, \cite{C98})
reported the presence of a strong optical modulation and attributed
this to an eclipse; the orbital period was subsequently determined to
be 6.0-hr both photometrically (Tomsick et al.\ 1998b, \cite{T98b};
Ilovaisky \& Chevalier 1998, \cite{IC98}) and spectroscopically (Hynes
et al.\ 1998, \cite{H98}). Tomsick et al.\ (1998b, \cite{T98b})
suggested that the 0.9-mag modulation is likely actually due to the
changing aspect of the heated companion in a high inclination system,
although partial eclipses appear also to be superposed on this
(Zurita, Casares \& Hynes 1998, \cite{ZCH98}). In this paper we
present the results of our spectrophotometric study of XTE~J2123--058
using the William Herschel Telescope (WHT), La Palma. Our photometric
observations are described in a companion paper in this proceedings,
Zurita et al.\ (1999, hereafter Paper I, \cite{Z99}).
\section{Our dataset}
We observed XTE J2123--058 through two 6-hr binary orbits on 1998 July
19--20. We used the blue arm of the ISIS dual-beam spectrograph on
the 4.2-m William Herschel Telescope to obtain 28 spectra. The R300B
grating combined with an EEV $4096\times2048$ CCD gave an unvignetted
coverage of $\sim$4000--6500\,\AA\ with some useful data outside this
range. An 0.7--1.0\,arcsec slit gave a spectral resolution
2.9--4.1\,\AA. Each spectrum was calibrated relative to a second star
on the slit. Absolute calibration was tied to the spectrophotometric
standard Feige 110 (Oke 1990, \cite{O90}). Wavelength calibration was
obtained from a copper-argon arc lamp, with spectrograph flexure
corrected using sky emission lines.
Our average spectrum shown in Fig.\ \ref{SpecFig} is derived from a
straight sum of count rates before slit loss and extinction
corrections to maximize the signal to noise ratio. The
spectral energy distribution was determined from an
average of calibrated spectra interpolated onto a uniform phase grid,
i.e.\ it is a uniformly weighted average over all phases.
\begin{figure}
\centering
\psfig{file=15_RHYNES1_1.ps,angle=90,width=12cm}
\caption{Average optical spectrum of XTE J2123--058.}
\label{SpecFig}
\end{figure}
\section{The line spectrum}
At first glance, XTE J2123--058 presents a nearly featureless blue
spectrum, with only the Bowen blend (N\,\textsc{iii}/C\,\textsc{iii}
4640\,\AA) and He\,\textsc{ii} 4686\,\AA\ prominent. In addition,
however, a number of weaker emission lines are present, the Balmer
lines exhibit complex profiles and weak interstellar absorption
features are seen.
\begin{figure}
\centering
\noindent\psfig{file=15_RHYNES1_2.ps,angle=90,width=12cm}
\centering
\noindent\psfig{file=15_RHYNES1_3.ps,angle=90,width=12cm}
\caption{Average optical spectrum of XTE J2123--058 in more detail.
All identified absorption or emission lines are marked.}
\label{LineSpecFig}
\end{figure}
\begin{table}[t]
\begin{center}
\begin{tabular}{rlll}
\hline
\noalign{\smallskip}
& Identification & Wavelength (\AA) & Comment \\
\noalign{\smallskip}
\hline
\noalign{\smallskip}
$\star$ & Ca\,\textsc{ii}\,K & 3933.7 &
Interstellar\\
& He\,\textsc{ii}\ Br$\theta$ & 4100.0 &
Blended with H$\delta$?\\
& H$\delta$ & 4101.7 & \\
& He\,\textsc{ii}\ Br$\eta$ & 4199.8 & \\
& Ca\,\textsc{ii}\ & 4220.1 & \\
& He\,\textsc{ii}\ Br$\zeta$ & 4338.7 &
Blended with H$\gamma$?\\
& H$\gamma$ & 4340.5 & \\
& DIB & 4428 &
Interstellar\\
$\star$ & He\,\textsc{ii}\ Br$\epsilon$ & 4541.6 & \\
$\star$ & N\,\textsc{iii}/C\,\textsc{iii} & 4640 & \\
$\star$ & He\,\textsc{ii}\ P$\alpha$ & 4685.7 & \\
& He\,\textsc{ii}\ Br$\delta$ & 4859.3 &
Blended with H$\beta$?\\
& H$\beta$ & 4861.3 & \\
$\star$ & He\,\textsc{ii}\ Br$\gamma$ & 5411.5 & \\
& C\,\textsc{iv} & 5801.5,5812.1 & \\
$\star$ & Na\,\textsc{i}\,D & 5890.0,5895.9 &
Interstellar\\
& He\,\textsc{ii}\ Br$\beta$ & 6560.1 &
Blended with H$\alpha$?\\
& H$\alpha$ & 6562.5 &
\end{tabular}
\vspace{3mm}
\caption{Spectral lines detected in XTE~J2123--058; $\star$ indicates a
definite detection.}
\label{EmissionTable}
\end{center}
\end{table}
Balmer lines from H$\beta$ to H$\delta$ appear to show broad
absorption and an emission core. The wavelength range marked
underneath each Balmer line in Fig.\ \ref{LineSpecFig} corresponds to
$\pm1500$\,km\,s$^{-1}$; this is intended to be an approximate guide
to the width, rather than a fit. The emission core may partly be Balmer
emission but at least some is attributable to coincident
He\,\textsc{ii} lines (since we see He\,\textsc{ii} 4542\,\AA\ and
5412\,\AA\, we would expect to also see related lines such as
4859\,\AA). This broad absorption plus narrow emission Balmer line
structure is also seen in other systems, for example the neutron star
LMXB 4U~2129+47 (Thorstensen \& Charles 1982, \cite{TC82}) and the
black hole candidate GRO~J0422+32 in outburst (Shrader et al.\ 1994,
\cite{S94}).
Based on the strength of the Na\,D1 line and the calibration of Munari
\& Zwitter (1997, \cite{MZ97}) we estimate $E(B-V) = 0.12 \pm 0.05$.
\section{Emission line behavior}
Both N\,\textsc{iii}/C\,\textsc{iii} 4640\,\AA\ (Bowen blend) and
He\,\textsc{ii} 4686\,\AA\ emission lines show changes in integrated flux
over an orbital cycle and He\,\textsc{ii} also reveals complex line
profile changes, with multiple S-wave components present in the
trailed spectrogram shown in Fig.\ \ref{SpectrogramFig}.
\begin{figure}
\centering
\psfig{file=15_RHYNES1_4.eps,height=6.2cm}
\hspace{\fill}
\psfig{file=15_RHYNES1_5.eps,height=6.2cm}
\caption{a) Emission line trailed spectrograms of
C\,\textsc{iii}/N\,\textsc{iii} 4640\,\AA\ and He\,\textsc{ii}
4686\,\AA\ in wavelength. b) Expansion of the He\,\textsc{ii}
spectrogram in velocity.}
\label{SpectrogramFig}
\end{figure}
The light curves of the two lines shown in Fig.\ \ref{LineLCFig}
appear somewhat different in structure; the Bowen blend peak is
broader and earlier than that of He\,\textsc{ii}. This suggests an
origin from different sites within the system. The Bowen blend light
curve in fact appears similar to the continuum light curves shown in Fig.\
\ref{LCFig}; Bowen emission may therefore originate on the heated
face of the companion star, with the modulation arising from the
varying visibility of the heated region.
\begin{figure}
\centering
\psfig{file=15_RHYNES1_6.ps,width=12cm}
\caption{Emission line light curves of N\,\textsc{iii}/C\,\textsc{iii}
and He\,\textsc{ii}.}
\label{LineLCFig}
\end{figure}
He\,\textsc{ii} emission, shows
a strong peak near phase 0.75 with a suggestion of a weaker one near
0.25. The modulation probably indicates that the emission region is
optically thick. We should be cautious in interpreting the light
curve, however, as the complex line behavior indicates multiple
emission sites. The integrated light curve is an average of different
light curves of several regions. A possible resolution to this
problem will be to use Doppler tomography to locate the dominant
emission sites (perhaps the two brightest spots). Then we can fit a
toy model in which each of these spots (treated as a point source) is
allowed to vary smoothly in brightness over an orbital cycle. This
procedure may allow us to approximately allow us to deconvolve the
light curves of the different regions.
\section{Doppler tomography}
We have used the technique of Doppler Tomography (Marsh \& Horne 1988,
\cite{MH88}) to identify emission sites in velocity space. Both the
back-projection method implemented in \textsc{molly} and
maximum-entropy method of \textsc{doppler} give similar results, as
does the alternative maximum entropy implementation of Spruit
\cite{S98}. Fig.\ \ref{TomogramFig} shows a maximum entropy tomogram
generated with \textsc{doppler}. Appropriate values were chosen for
instrumental resolution ($\sim180$\,km\,s$^{-1}$) and phase smearing.
We overplot the position of the Roche lobe of the companion, the
accretion stream ballistic velocity, the Keplerian velocity along the
accretion stream and the Keplerian velocity around the disk
edge. These are derived from uncertain system parameters, so they
should be viewed cautiously. The parameters are determined from
light curve fits (see Paper I). We should also beware that one of the
fundamental assumptions of Doppler tomography, that we always see all
of the line flux at {\it some} velocity, is clearly violated, as the
integrated line flux is not constant.
\begin{figure}
\centering
\psfig{file=15_RHYNES1_7.ps,width=12cm}
\caption{Maximum entropy He\,\textsc{ii} 4686\,\AA\ tomogram.
Overplotted are the center-of-mass ($\oplus$), the centers of the two
stars ($+$), the companion star Roche lobe, the ballistic accretion
stream velocity (solid), the Keplerian velocity along the stream
position (dotted) and the Keplerian velocity at the disk edge
(circle). All of these are based on the parameters of Paper I:
$M_1=1.8\,M_{\odot}$, $M_2=0.6\,M_{\odot}$, $T=0.24821$\,d,
$i=74^{\circ}$, $r_{\rm disk}=0.73\,r_{\rm L1}$.}
\label{TomogramFig}
\end{figure}
In spite of these cautions, we can learn something from the exercise.
The dominant emission site (corresponding to the main S-wave) appears
on the opposite side of the neutron star from the companion. It is
inconsistent with the heated face of the companion and the stream/disk
impact point, although a tail does appear to extend upwards towards
the expected stream position. As the emission appears to form an arc
roughly centered on the neutron star position, it is tempting to
associate it with asymmetric disk emission. Unfortunately the
velocity of the strongest emission is too low for disk material. This
can be seen from the fact that it lies inside the circle representing
the Keplerian velocity at the disk edge; the inner disk will have {\em
higher} velocities than this. If the observed bright spot is indeed
emission from the disk then it must come from sub-Keplerian material.
A more promising explanation is suggested by the similarity to some SW
Sex type cataclysmic variables (e.g.\ V1315 Aql; compare H$\beta$
tomograms in Dhillon, Marsh and Jones 1991, \cite{DMJ91}, and Hellier
1996, \cite{H96}). This is that the emission is actually associated with an
extension of the accretion stream beyond its nominal disk impact
point. One possible model involves a disk-anchored magnetic
propeller (Horne 1999 \cite{H99}), suggested for SW Sex systems, which
ejects some of the stream material from the system. An alternative is
that some material splashes from the stream impact point, rising high
above the disk. Such material will follow a trajectory similar to
that seen, with the brightest observed spot corresponding to the point
where this splashing material reimpacts the disk.
\section{The spectral energy distribution}
We show in Fig.\ \ref{SEDFig} our average spectrum after dereddening
using the extinction curve of Cardelli et al.\ 1989, \cite{CCM89}) and
our reddening estimate of $\rm E(B-V)=0.12$ derived from the Na\,D1
line (see above). We also show the photometry of Tomsick et al.\
(1998a, \cite{T98a}) from 1998 June 30 and a spectrum of GRO~J0422+32
(1992 August 17) provided by C.R. Shrader. This black hole X-ray
transient has a 5.1-h orbital period, making it the black hole system
most similar to XTE~J2123--058. Both the photometry
of Tomsick et al. and our spectroscopy appear steeper than the
spectrum of GRO~J0422+32, taking a steep blue power-law form. In view
of the difficulties of accurately calibrating U band data, it is
unclear whether the apparent flattening off of the spectrum at high
energies is real or an artifact.
\begin{figure}
\centering
\psfig{file=15_RHYNES1_8.ps,angle=90,width=12cm}
\caption{Dereddened spectral energy distributions of XTE~J2123--058
and the black-hole X-ray transient GRO~J0422+32. Also plotted are the
photometry of Tomsick et al.\ (1998a, \cite{T98a}) and approximately
fitting power-law spectra for comparison.}
\label{SEDFig}
\end{figure}
\section{Continuum behavior}
We show in Fig.\ \ref{LCFig} light curves for three `continuum' bins
at 4500\,\AA, 5300\,\AA\ and 6100\,\AA. The light curves show very
similar shapes, with no significant differences in profile or
amplitude within this wavelength range. The apparent differences
between them, most noticeable around phase 0.6, are likely due to
calibration uncertainties: coverage on each night was approximately
from phase 0.6 through 1.6.
\begin{figure}
\centering
\psfig{file=15_RHYNES1_9.ps,width=12cm}
\caption{Continuum light curves. Formal errors are smaller than the
points, but systematic errors are larger, and are probably responsible
for the apparent differences between wavelengths.}
\label{LCFig}
\end{figure}
The light curve morphology is fully discussed in Paper I. We believe
it is mainly due to the changing aspect of the heated companion star,
which is the dominant light source at maximum light, near phase 0.5.
At minimum light (phase 0.0) the heated face is obscured and we see
the accretion disk only. At all phases the unilluminated parts of the
companion star are expected to contribute negligible flux as the
outburst amplitude is $\sim 5$ magnitudes.
The lack of strong color dependence then indicates that either the
disk and heated companion have similar temperatures, or that both are
sufficiently hot that we only see the $F_{\nu} \propto \nu^{2}$
Rayleigh-Jeans part of the spectrum. The very steep spectral energy
distribution shown in Fig.\ \ref{SEDFig} suggests that the latter is
the case.
\section{Conclusion}
Our main findings are summarized below. Where appropriate, we make
comparisons with the black hole X-ray transient, GRO~J0422+32. This
system has a 5.1-h orbital period and may be the black hole system
most similar to XTE~J2123--058. We also note some similarities to the
neutron star LMXB 4U~2129+47 which has a 5.2-hr orbital period and
shows similar large amplitude photometric variations to XTE~J2123--058.
\begin{itemize}
\item
High excitation emission lines dominate the spectrum (He\,\textsc{ii},
N\,\textsc{iii}/C\,\textsc{iii}, C\,\textsc{iv} (see upper left
panel). No He\,\textsc{i} emission is seen. The blue line spectrum
looks rather similar to that of 4U~2129+47 (Thorstensen \& Charles
1982, \cite{TC82}).
\item
He\,\textsc{ii} emission is dominated by a region that coincides with
neither the heated companion star, the ballistic accretion stream nor
a Keplerian disk. This is unlike GRO~J0422+32 (Casares et al.\ 1995,
\cite{C95}), the only other transient for which outburst Doppler
tomography has been performed. In GRO~J0422+32, He\,\textsc{ii}
emission appears to originate from the accretion stream/disk impact
point. Our observations may possibly be explained by extension of the
stream beyond the initial impact point. This may be similar to the
He\,\textsc{ii} emission in 4U~2129+47 (Thorstensen \& Charles 1982,
\cite{TC82}) which has a radial velocity modulation with similar phase
and amplitude to that we see.
\item
The continuum spectral energy distribution is very blue, and steeper
than in black hole X-ray transients such as GRO~J0422+32 (see Fig.\
\ref{SEDFig}). This implies that both the heated face of the
companion star, and the disk, are hotter than in similar short-period
black hole systems. Such a conclusion is supported by the dominance
of high-excitation emission lines. We suggest tentatively that this
may indicate more efficient X-ray irradiation in this (neutron star)
transient compared to black hole systems. This has previously been
suggested by King, Kolb and Szuszkiewicz (1997, \cite{KKS97}) as an
explanation for why most transient LMXBs contain black holes whereas
apparently all persistent sources contain neutron stars. Before we
can claim this is a firm conclusion we will need to assess the
uncertainty in the spectral energy distribution and perform a more
systematic comparison with other systems taking into account
differences in X-ray luminosity.
\end{itemize}
\section*{Acknowledgments}
Thanks to Chris Shrader for providing the spectrum of GRO J0422+32 for
comparison. Doppler tomography used \textsc{molly} and
\textsc{doppler} software by Tom Marsh and \textsc{dopmap} software by
H.C. Spruit (\cite{S98}). The William Herschel Telescope is operated
on the island of La Palma by the Isaac Newton Group in the Spanish
Observatorio del Roque de los Muchachos of the Instituto de
Astrofisica de Canarias. RIH is supported by a PPARC Research
Studentship.
\section*{References}
|
\section{General properties of fibered Calabi-Yau threefolds}
\subsection{Fiber spaces }
Let $\pi:X\longrightarrow} \def\sura{\twoheadrightarrow Y$ be an algebraic fiber space, i.e., a
proper morphism of algebraic varieties with connected fibers; in general
we will be assuming the base $Y$ of the fibration is smooth and the fibers
are generically smooth. The {\it discriminant} $\Delta\subset} \def\nni{\supset} \def\und{\underline Y$ is the locus of
$y\in Y$ such that the fiber $X_y$ over $y$ is not smooth, in other words
the image of the set of points for which $d\pi$ fails to have maximal
rank. Assume that $X$ is also smooth, and let $K_X,\ K_Y$ denote the
canonical bundles. The {\it relative canonical bundle} is $K_{X|Y}$,
defined as
\begin{equation}\label{1} K_{X|Y}=\pi^*K_Y^{-1}\otimes K_X.
\end{equation}
It follows from (\ref{1}) that $\pi_*K_{X|Y} =K_Y^{-1}\otimes \pi_*K_X$, so
in particular
\begin{equation}\label{2} K_X=\cO_X \Ra \pi_*(K_X)=\cO_Y \Ra
\pi_*K_{X|Y} = K_Y^{-1}.
\end{equation}
This formula gives a {\it necessary} condition on a fiber space $\pi:X\longrightarrow} \def\sura{\twoheadrightarrow
Y$ for $X$ to be Calabi-Yau. Note also that equation (\ref{2}) is an
equation for {\it divisors} on the base $Y$ of the fibration.
\subsection{Fiber spaces with section }
Let $\pi:X\longrightarrow} \def\sura{\twoheadrightarrow Y$ be an algebraic fiber space, and assume
now that $\pi$ has a {\it section} $\sigma:Y\longrightarrow} \def\sura{\twoheadrightarrow X$. Let $\Sigma=\sigma(Y)$; we can
identify $Y$ with the subvariety $\Sigma$ of $X$. As such $\Sigma$ has a normal
bundle $N_X\Sigma$, and we may apply adjuction, yielding (now as an equation
of divisors on $\Sigma$, written additively)
\begin{equation}\label{3} {K_X}_{|\Sigma} = K_{\Sigma} - c_1(N_X\Sigma),
\end{equation}
where $c_1$ denotes the first Chern class, an equivalence class of
divisors. Thus the necessary condition (\ref{2}) takes the form:
\begin{equation}\label{4} c_1(N_X\Sigma)=K_{\Sigma}.
\end{equation}
\subsection{Elliptic fibrations }
Assume now that $X\longrightarrow} \def\sura{\twoheadrightarrow Y$ is elliptic with a section
$\sigma:Y\longrightarrow} \def\sura{\twoheadrightarrow X$ and let as above $\Sigma$ denote the image $\Sigma=\sigma(Y)$. It
follows from fundamental results of Nakayama \cite{Na}, 2.1, that there is
a Weierstra\ss\ model $W(\cL,g_2,g_3)$ over $Y$ and a proper birational map
$\mu:X\longrightarrow} \def\sura{\twoheadrightarrow W(\cL,g_2,g_3)$ over $Y$ such that $\mu(\Sigma)$ is the {\it zero
section} of the Weierstra\ss\ model. A Weierstra\ss\ model over $Y$
consists of (a) a line bundle $\cL$ over $Y$, and (b) sections $g_2\in
{\rm H}^0(Y,\cL^{\otimes -4}),\ g_3\in {\rm H}^0(Y,\cL^{\otimes -6})$. The discriminant
$\Delta:= g_2^3-27 g_3^2 \in {\rm H}^0(Y,\cL^{\otimes -12})$ is the discriminant of
the Weierstra\ss\ model $W(\cL, g_2,g_3)$, which is defined as follows. Fix
meromorphic sections $x, y, z$ of $\cO_{\fP}(1)\otimes \cL^{-2},\
\cO_{\fP}(1)\otimes \cL^{-3}$ and $\cO_{\fP}(1)$, respectively, where
$\fP:=Proj(\cO_Y\oplus \cL^2\oplus \cL^3)$. Then $W(\cL,g_2,g_3)$ is the
divisor on $\fP$ defined by
\begin{equation}\label{5} y^2 z = 4x^3 +g_2 x z^2 + g_3 z^3.
\end{equation}
The {\it zero section} of this Weierstra\ss\ model is determined by
dehomoginizing, i.e., choosing the inflection point at infinity of the
Weierstra\ss\ cubic as the zero point of the curve. Note that, if $\Sigma$
denotes the image of a section as above, then $\cL \cong N_X\Sigma$, where we
view $\cL$ as a line bundle on $\Sigma$. Then it follows that $\Delta = -12
c_1(N_X\Sigma)$.
A Weierstra\ss\ model
is {\it minimal}, if there is no prime divisor $D$ on $Y$ such that
$div(g_2)\geq 4D$ and $div(g_3)\geq 6D$. Nakayama shows that if $Y$ is
smooth and the discriminant $\Delta$ has normal crossings, then $W$ has only
rational singularities, if and only if, $W(\cL,g_2,g_3)$ is
minimal.
\begin{lemma} Let $X\longrightarrow} \def\sura{\twoheadrightarrow Y$ be elliptic with section. The condition
(\ref{4}) is necessary and sufficient for $X$ to be Calabi-Yau.
\end{lemma}
{\bf Proof:} We must show that $K_{X|\Sigma}=\cO_{\Sigma}$ implies
$K_X=\cO_X$. This follows from Kodaira's formula for the canonical bundle,
in the higher-dimensional formulation as in \cite{Ka}, which is\footnote{in
general there will be an additional error term which is present due to
less accurate control over the birational geometry of these spaces in
higher dimensions; by results of Grassi \cite{G} this term can be avoided
by choosing our model correctly}
\begin{equation}\label{6} K_X = \pi^*(K_Y + \sum a_i[\Delta_i]),
\end{equation}
where $\Delta_i$ are the irreducible components of the discriminant $\Delta$ and
the rational numbers $a_i$ are determined by the type of singular fiber
over $\Delta_i$. For every $a_i$ we have $12 a_i \in {\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}$, and setting $\Delta =
\sum 12 a_i \Delta_i$, the divisor $\Delta$ is divisible by 12 with $\Delta = \sum
12 a_i \Delta_i = -12 c_1(N_X\Sigma) = -12c_1(\cL) \Ra c_1(\cL)=-\sum a_i \Delta_i$. At
any rate, (\ref{6}), (\ref{4}) and the assumption $K_{X|\Sigma} = \cO_{\Sigma}$
together give $K_X = \pi^*(K_{X|\Sigma}) = \pi^*(\cO_{\Sigma}) = \cO_X$. \hfill $\Box$ \vskip0.25cm } \def\cD{\ifmmode {\cal D
Finally, if $X\longrightarrow} \def\sura{\twoheadrightarrow Y$ is an elliptic fibration and $X$ is a smooth threefold
with $K_X=\cO_X$, then $X$ is birational to a minimal Weierstra\ss\ model
over a surface $S$ which is one of the following:
\begin{enumerate}\item a minimal surface with $\kappa(Y)=0$, $g_2,\ g_3$
constant. \item a minimal ruled surface over an elliptic curve. \item
$S=\fP^2$ or $S$ is one of the Hirzebruch surfaces $\fF_n,\ 0\leq n \leq 12$.
\end{enumerate}
(For $n> 12$, any Weierstra\ss\ model is necessarily non-minimal, hence by
the above result has non-rational singularities and is consequently not
Calabi-Yau).
\subsection{Abelian surface fibrations }
Next we assume that the general fiber of $X\longrightarrow} \def\sura{\twoheadrightarrow Y$ is an
abelian surface, and let $\sigma:Y\longrightarrow} \def\sura{\twoheadrightarrow X$ be a section. Now the image $\Sigma$
has codimension two, being just a point in each fiber. The normal bundle
$N_X\Sigma$ is a rank two bundle on $\Sigma$ (and on $Y$), and $c_1(N_X\Sigma)$ is a
divisor on $\Sigma$, so the formula (\ref{4}) still makes sense. Here we also
have a formula for the canonical bundle \cite{A}, 2.16, which is the
following:
\begin{equation}\label{7} K_X = \pi^*(K_Y + \sum a_i[\Delta_i]) +
\sum_i^N(e_jF_j + E_j) + \sum_1^M (d_iD_i + D_i'),
\end{equation}
where the first sum is over all singular fibers, the second sum over all
{\it multiple} fibers and the third sum is over all components of the
singular fibers which contain {\it two or more} components. It follows from
this that we have
\begin{lemma} Let $X\longrightarrow} \def\sura{\twoheadrightarrow Y$ be a fiber space of abelian varieties over a
curve, and suppose that
\begin{enumerate} \item There are no multiple fibers. \item All singular
fibers are irreducible.
\end{enumerate}
Then the condition (\ref{4}) is necessary and sufficient for $X$ to be
Calabi-Yau.
\end{lemma}
{\bf Proof:} Using the formula (\ref{7}), the proof is just as
above, the important fact being that under the stated assumptions $K_X$ is
the pullback of a class on the base. \hfill $\Box$ \vskip0.25cm } \def\cD{\ifmmode {\cal D
\subsection{K3-fibrations}
In the case that $X$ is a K3-fibration, assuming $X$ is a Calabi-Yau
threefold, the base $Y$ is necessarily $\fP^1$. The necessary condition
(\ref{4}) becomes in this case $c_1(N_X\Sigma) =-2$. As we shall see below,
however, this condition is {\it not} sufficient, not even under assumptions
as in the previous lemma. We will give an example of a K3-fibration, for
which (\ref{4}) is satisfied, without multiple fibers, and for which all
singular fibers are irreducible, but which is far from being Calabi-Yau. To
discuss this example, we first discuss monodromy and how this relates to
the coefficients $a_i$ which occur in (\ref{6}) and (\ref{7}).
\section{Torsion monodromy and fibrations with constant modulus}
\subsection{Monodromy}
Let $X\longrightarrow} \def\sura{\twoheadrightarrow Y$ be an algebraic fiber space, and let $Y_0\subset} \def\nni{\supset} \def\und{\underline Y$ denote the
Zariski open subset of $Y$ over which $\pi$ is smooth, i.e., the complement
of the discriminant $\Delta$.
Let $\pi_0:X_0\longrightarrow} \def\sura{\twoheadrightarrow Y_0$ denote the
restriction of $\pi$ to $X_0=X-\pi^{-1}(Y-Y_0)$, so that $\pi_0$ {\it is} a
locally trivial fibration in the sense of topology. The collection of the
integral homology groups of the fibers, ${\rm H}^k(F_y,{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P})$, ($y\in Y_0$) form a
sheaf $
\cF_0$ over $Y_0$, denoted $R^k(\pi_0)_*{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}$, and the
monodromy around any $s\in \Delta:=Y-Y_0$ can be viewed as a translation in
the fiber over some fixed base point $*\in Y_0$, $(\cF_0)_*$; sending
$\gamma \in \pi_1(Y_0,*)$ to the matrix $T_{\gamma}\in \hbox{Aut}} \def\Im{\hbox{Im}} \def\mod{\hbox{mod}((\cF_0)_*) =
\hbox{Aut}} \def\Im{\hbox{Im}} \def\mod{\hbox{mod}({\rm H}^k(F_*,{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}))$ which describes this translation gives the monodromy
representation \[ \rho:\pi_1(Y_0,*) \longrightarrow} \def\sura{\twoheadrightarrow \hbox{Aut}} \def\Im{\hbox{Im}} \def\mod{\hbox{mod}((\cF_0)_*).\]
By the monodromy theorem, each monodromy matrix around an isolated bad
fiber has eigenvalues which are roots of unity. Let $T$ be the monodromy
matrix; if $(T-1)$ is {\it nilpotent}, one says $T$ is {\it
unipotent}. Suppose $T^k=1$ for some $k$; in this case we will speak of
{\it torsion} monodromy. The general monodromy matrix is {\it
quasi-unipotent}, meaning that for some $N$ and $k$ we have
$(T^k-1)^N=0$. In this case, $N$ is called the degree of unipotency, and it
is at most the dimension of the fiber.
Suppose that we have a {\it local} fiber space $\cX \longrightarrow} \def\sura{\twoheadrightarrow D$ over the unit
disc $D$, and assume that all fibers $\cX_z$ are smooth for $z\in D^*=D -
\{ 0 \}$. Assume furthermore that the monodromy is torsion, say $T^k=1$,
and consider the base change $D\longrightarrow} \def\sura{\twoheadrightarrow D,\ z\mapsto z^k$. Pulling back the
fibration $\cX$, we see that the monodromy is now $T^k=1$, i.e.,
trivial. Again, if $X\longrightarrow} \def\sura{\twoheadrightarrow Y$ is an algebraic fiber space over a curve $Y$,
and all monodromies are torsion, then we can find a (branched) cover
$Y'\longrightarrow} \def\sura{\twoheadrightarrow Y$, such that the monodromy of the pull-back of $\cX$ is
trivial. For this it is sufficient to have the branching of order $k_i$ at
each point $z_i$ for which the local monodromy matrix $T_i$ has order
$k_i$.
While the sheaf $\cF_0$ is a sheaf of ${\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}$-modules, by tensoring with
$\cO_{Y_0}$ we get a sheaf of $\cO_{Y_0}$-modules, call it
\[{\cal E}} \def\cF{{\cal F}} \def\fF{{\bf F}_0:=\cF_0\otimes \cO_{Y_0}.\]
There is a very special extension of ${\cal E}} \def\cF{{\cal F}} \def\fF{{\bf F}_0$, called the {\it canonical
extension} (Schmidt) ${\cal E}} \def\cF{{\cal F}} \def\fF{{\bf F}$ of ${\cal E}} \def\cF{{\cal F}} \def\fF{{\bf F}_0$ to $Y$, i.e., such that
${\cal E}} \def\cF{{\cal F}} \def\fF{{\bf F}_{|Y_0}={\cal E}} \def\cF{{\cal F}} \def\fF{{\bf F}_0$. It is known how to describe holomorphic sections
generating ${\cal E}} \def\cF{{\cal F}} \def\fF{{\bf F}$: let $r=rk_{{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}}{\rm H}^k(F,{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P})$, and let $v_1,\ldots, v_r$ be
a ${\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}$-basis. For each $s\in \Delta$ let $T=T_s$ be the monodromy map
defined by a small loop around $s$, and let $t=t_s$ be a local coordinate on
the base near $s$ (i.e., $t=0$ defines the point $s\in Y$). Then, for
$j=1,\ldots,r$, the expressions
\[ \sigma_j=\exp\left({1\over 2\pi i}\log T\log t\right) v_j\]
define holomorphic sections of ${\cal E}} \def\cF{{\cal F}} \def\fF{{\bf F}$ and in fact generate it. If this
expression seems somewhat formidable\footnote{The first named author found
it so; he is indebted to Donu Arapura for the following explanation},
consider the case that we only have
torsion monodromy.
Consider a Galois cover of $Y$ such that for each $s_i\in \Delta$, if $k_i$
denotes the order of $T_{s_i}$ ($T_{s_i}^{k_i}=1$), then the cover
$p:Y'\longrightarrow} \def\sura{\twoheadrightarrow Y$ is given locally by $t_{s'}'\mapsto (t_{s'}')^{k_i}=t_s$ for
any $s'\in p^{-1}(t_{s_i}),\ s_i\in \Delta$. By the results above, the lift of
$X\longrightarrow} \def\sura{\twoheadrightarrow Y$ to $Y'$,
\[\begin{array}{rcccl} & X' & \stackrel{p}{\longrightarrow} \def\sura{\twoheadrightarrow} & X & \\
\pi' & \downarrow & & \downarrow & \pi \\
& Y' & \stackrel{p}{\longrightarrow} \def\sura{\twoheadrightarrow} & Y &
\end{array}\]
will have trivial monodromy. The eigenvalues of $T$ are of the form
$e^{2\pi i \alpha_j},\ j=1,\ldots, r$, where $\alpha_j=p_j/k_i$ are
rational numbers with $k_i$, $p_j\in {\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}$ (not necessarily in
lowest terms). Let us suppose that $T$ is diagonalized;
then the matrix $ \log T$ can be calculated as follows:
\[ \log T = \log \left(\begin{array}{ccc} e^{2\pi i \alpha_1} & & 0 \\ &
\ddots & \\ 0 & & e^{2\pi i \alpha_r}
\end{array}\right) = \left(\begin{array}{ccc} 2\pi i \alpha_1 & & 0 \\ &
\ddots & \\ 0 & & 2\pi i \alpha_r
\end{array}\right),\]
so that the expression $ \exp\left({1\over 2\pi i}\log T\log
t\right)$ can be calculated as
\[ \exp\left({1\over 2\pi i}\log T\log
t\right) =\left(e^{\log t}\right)^{{1\over 2\pi i}\log T} =
t^{\left({1\over 2\pi i} \log T\right)} = \left(\begin{array}{ccc}
t^{\alpha_1} & & 0 \\ & \ddots & \\ 0 & & t^{\alpha_r}
\end{array}\right), \]
and we have $\sigma_j= \left(\begin{array}{ccc}
t^{\alpha_1} & & 0 \\ & \ddots & \\ 0 & & t^{\alpha_r}
\end{array}\right) v_j$ for $j=1,\ldots, r$. So in this case we see how
to calculate generators of ${\cal E}} \def\cF{{\cal F}} \def\fF{{\bf F}$ -- they are just certain (rational) powers
of the local coordinate times the locally constant section.
Now consider the sheaves $\pi_*K_{X|Y}$ and $\pi'_* K_{X'|Y'}$.
This sheaf is related to the canonical extension $
{\cal E}} \def\cF{{\cal F}} \def\fF{{\bf F}$ above by the following result of Kawamata (\cite{Ka2}, Theorem 1):
\[ \pi_* K_{X|Y} \cong i_* {\cal E}} \def\cF{{\cal F}} \def\fF{{\bf F}_0^{k,0} \cap {\cal E}} \def\cF{{\cal F}} \def\fF{{\bf F},\]
where $i:Y_0\rightarrow Y$ is the inclusion and ${\cal E}} \def\cF{{\cal F}} \def\fF{{\bf F}_0^{k,0}\subset} \def\nni{\supset} \def\und{\underline {\cal E}} \def\cF{{\cal F}} \def\fF{{\bf F}_0$ is
the ${\rm H}^{k,0}$-part of the Hodge decomposition of ${\rm H}^k(F,\fC)$. Note this is
an isomorphism of line bundles, and the monodromy group acts here with a
single root of unity at each $\Delta_i\subset} \def\nni{\supset} \def\und{\underline \Delta$; this root of unity
$e^{2\pi i \alpha_i}$ gives the coefficient of $\Delta_i$ in the sheaf
$\pi_* K_{X|Y}$ above.
We now apply this to K3 fibrations $X\longrightarrow} \def\sura{\twoheadrightarrow \fP^1$, the $\Delta_i$ are points
on $\fP^1$, $k=2$, and the result above shows that:
\begin{quote} the action of the monodromy on the holomorphic two-form
determines the coefficient $r_i$ of $\Delta_i$ in $\pi_*K_{X|Y}$ by the
rule:
\begin{equation} \label{formula} \hbox{If } T_i(\omega) =
e^{2\pi i \alpha_i} \omega \quad \hbox{( $\omega$ the
holomorphic two-form)},
\end{equation}
then $r_i = \alpha_i$, i.e., $\pi_*K_{X|Y} = \sum \alpha_i \Delta_i$. In
particular, $\sum \alpha_i =2$ is a necessary condition for $X$ to be
Calabi-Yau.
\end{quote}
\subsection{Constant modulus} In general, given an algebraic fiber space
$X\longrightarrow} \def\sura{\twoheadrightarrow Y$, the complex structure (modulus) of the fibers vary, in a
holomorphic manner. However, it can also occur that the modulus is fixed; in
this case we speak of {\it constant modulus}. Easy examples are given by
elliptic surfaces for which the $J$-invariant is a constant; indeed,
suppose $X\longrightarrow} \def\sura{\twoheadrightarrow Y$ is an elliptic fibration, and let $W\longrightarrow} \def\sura{\twoheadrightarrow Y$ be the
Weierstra\ss\ model. Then $J=g_2^3 / \Delta$, and $J$ is constant when
$g_2$ or $g_3$ vanish. For example, suppose $g_2=0$, so that $J\equiv 0$. Then
all singular fibers are of types $II,\ II^*, IV, IV^*$. More precisely,
suppose we choose $Y=\fP^1$, and let $\Delta$ consist of two points, with
singular fibers of type $II$ at one and $II^*$ at the other. This is a
rational elliptic surface, and clearly all elliptic fibers have $J =0$ and
so are copies of
the elliptic curve with automorphism group ${\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}/6{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}$. In this case by the
result of the previous section, there is a cover $Y'\longrightarrow} \def\sura{\twoheadrightarrow \fP^1$, such that
the pull back of the fibration has trivial monodromy. Furthermore, the
modulus is still constant. Under these circumstances, it is clear that the
pull back of the fibration is a fibration without monodromy and with
constant modulus. It follows that this pull back is a {\it product}. This
fact holds more generally.
\subsection{Uniformisation} Let $X\longrightarrow} \def\sura{\twoheadrightarrow Y$ be an algebraic fiber space
satisfying the following conditions:
\begin{enumerate} \item The dimension of $Y$ is one.
\item All monodromies around bad fibers are
torsion. \item The modulus of the fibers of $X$ is constant.
\end{enumerate}
\begin{lemma} Under the assumptions just made, there is a finite, branched
cover $Y'\longrightarrow} \def\sura{\twoheadrightarrow Y$ such that the pull back of $X$ to $Y'$ is birational to
a product.
\end{lemma}
{\bf Proof:} For this it is sufficient to construct a cover $Y'\longrightarrow} \def\sura{\twoheadrightarrow Y$,
which is branched at each of the base points $y_i\in Y$
of the bad fibers to degree
$k_i$, where $T_i^{k_i}=1$, $T_i$ the local monodromy matrix at the point
$y_i$. Since the dimension of $Y$ is one, this can clearly be done. \hfill $\Box$ \vskip0.25cm } \def\cD{\ifmmode {\cal D
\noindent{\bf Remark:} Under some mild assumptions of the moduli space of
the fiber in question (as in, for example, \cite{jraw}), the assumption 3.
{\it implies} the assumption 2.\ above. If this is the case and the general
fiber is smooth, the assumption 3.\ means that the period map has image
a point which is in the {\it interior} of the moduli space. Since the
corresponding moduli point is for a smooth variety, the only degeneration
which can occur is a torsion one, gotten essentially by taking a finite
quotient of a smooth fiber. Thus 3.\ $\Ra$ 2. \hfill $\Box$ \vskip0.25cm } \def\cD{\ifmmode {\cal D
\section{A construction of quotients as weighted projective hypersurfaces}
\subsection{Notations}
We will be working with weighted projective spaces, which are certain
(singular) quotients of usual projective space. Alternatively, they may be
described as quotients of $\fC^{n+1}$ by a $\fC^*$-action. We assume the
weights $(w_0,\ldots, w_n)$ are given, let ${\bgm}_{w_i}$ denote the group
of $w_i$th roots of unity, and consider the action of ${\bgm}:=
{\bgm}_{w_1}\times \cdots \times {\bgm}_{w_n}$ on $\fP^n$ as follows. Let
$g=(g_0,\ldots, g_n) \in {\bgm}$, and consider for $(z_0:\ldots: z_n)$
homogenous coordinates on $\fP^n$ the action
\[ (g,(z_0:\ldots: z_n)) \mapsto (g_0z_0:\ldots: g_nz_n).\]
Alternatively, consider the action of $\fC^*$ on $\fC^{n+1}$ given by
\[ (t,(z_0,\ldots, z_n)) \mapsto (t^{w_0}z_0,\ldots, t^{w_n}z_n).\]
In both cases, the resulting quotient is the weighted projective space,
which we will denote by $\fP_{(w_0,\ldots, w_n)}$. General references for
weighted projective spaces are \cite{dolg} and \cite{Y}. A {\it weighted
hypersurface} is the zero locus of a weighted homogenous polynomial
$p$. Such a hypersurface or the corresponding polynomial is called {\it
transversal}, if the only singularities are the intersections with the
singular locus of the ambient weighted projective space, and quasismooth,
if the cone over the hypersurface is quasismooth, i.e., smooth outside of
the vertex; for weighted {\it hypersurfaces}, these notions are equivalent
(cf.\ \cite{Y}, Propositions 6 and 8).
We will assume the weights are {\it normalized} in the sense that no
$n$ of the $n+1$ weights have a common divisor $>1$. Both for the weighted
projective spaces as well as for the weighted hypersurfaces this assumption
is no restriction (cf.\ \cite{dolg} 1.3.1 and \cite{Y}, pp.\ 185-186).
We will write such isomorphisms in the sequel without further comment, for
example $\fP_{(2,3,6)}\cong \fP_{(2,1,2)}\cong \fP_{(1,1,1)}=\fP^2$, where
the first equality is because the last two weights are divisible by 3, the
second while the first and last are divisible by 2.
We will use the notation $\fP_{(w_0,\ldots, w_n)}[d]$ to denote either a
certain weighted hypersurface of degree $d$, or to denote the whole family
of such (the context will make the usage clear). In the particular case
that the weighted polynomial $p$ is of Fermat type, then there is a useful
fact, corresponding to the above normalizations. For example, in
$\fP_{(2,3,6)}$ consider the weighted hypersurface $x_0^6+x_1^4
+x_2^2=0$. Then the isomorphism $\fP_{(2,3,6)}\cong \fP_{(2,1,2)}$ above is
given by the introduction of new variables $(x_0')=x_0^3,$
which is in spite of appearances a one to one coordinate transformation
(becuase of admissible rescalings), and the Fermat polynomial becomes
$(x_0')^2+x_1^4+x_2^2=0$. Again, the isomorphism $\fP_{(2,1,2)}\cong
\fP_{(1,1,1)}$ is given by setting $(x_1')=x_1^2$, and the Fermat
polynomial becomes $(x_0')^2+(x_1')^2+x_2^2=0$, which is a quadric in the
projective plane. We {\it denote} this process by the symbolic expressions
\[ \fP_{(2,3,6)}[12]\cong \fP_{(2,1,2)}[4] \cong \fP_{(1,1,1)}[2].\]
\subsection{The construction}
We now introduce the twist map; this map will give an explicit form to the
forming of quotients of products $V_1\times V_2$ of weighted hypersurfaces
by an abelian group acting on the product.
\subsubsection{The twist map} Let $V_1,\ V_2$ be weighted hypersurfaces
defined as follows.
\begin{equation}\label{8} \parbox{10cm}{$\displaystyle V_1 = \{x_0^{\ell}
+p(x_1,\ldots,x_n) =0 \} \subset} \def\nni{\supset} \def\und{\underline \fP_{(w_0,w_1,\ldots, w_n)} \\ V_2 = \{
y_0^{\ell} + q(y_1,\ldots, y_m) =0 \} \subset} \def\nni{\supset} \def\und{\underline \fP_{(v_0,v_1,\ldots,v_m)}$},
\end{equation}
where we assume both $p$ and $q$ are quasi-smooth. The degrees of these
hypersurfaces are
\[ \nu=\hbox{deg}} \def\Pic{\hbox{Pic}} \def\Jac{\hbox{Jac}(V_1) = \ell\cdot w_0,\quad \mu=\hbox{deg}} \def\Pic{\hbox{Pic}} \def\Jac{\hbox{Jac}(V_2) = \ell\cdot v_0.\]
We then consider the hypersurface
\begin{equation}\label{9} X:= \{ p(z_1,\ldots, z_n) -q(t_1,\ldots, t_m) =0
\} \subset} \def\nni{\supset} \def\und{\underline \fP_{(v_0 w_1, \ldots, v_0 w_n, w_0 v_1, \ldots, w_0 v_m )}.
\end{equation}
Note that the degree of $X$ is $v_0\cdot \hbox{deg}} \def\Pic{\hbox{Pic}} \def\Jac{\hbox{Jac}(p)=w_0\cdot \hbox{deg}} \def\Pic{\hbox{Pic}} \def\Jac{\hbox{Jac}(q) =
v_0w_0\ell$.
\begin{lemma}\label{l3.1} The rational map
\begin{eqnarray*} \Phi: \fP_{(w_0,w_1,\ldots, w_n)} \times
\fP_{(v_0,v_1,\ldots,v_m)} & \longrightarrow} \def\sura{\twoheadrightarrow & \fP_{(v_0 w_1, \ldots, v_0 w_n, w_0
v_1, \ldots, w_0 v_m )} \\ ((x_0,\ldots, x_n),(y_0,\ldots, y_m)) &
\mapsto & (y_0^{w_1/w_0}\cdot x_1, \ldots, y_0^{w_n/w_0}\cdot x_n,
x_0^{v_1/v_0}\cdot y_1, \ldots, x_0^{v_m/v_0}\cdot y_m)
\end{eqnarray*}
restricts to $V_1\times V_2$ to give a rational generically finite
map onto $X$.
\end{lemma}
{\bf Proof:} First we show that the map is well-defined outside of the
locus\footnote{As the point $\{(1,0,\ldots, 0), (1,0,\ldots,0)\}$ is not
contained on $V_1\times V_2$ we will disregard this in what follows}
\[\{y_0=x_0=0\}\cup \{(1,0,\ldots, 0), (1,0,\ldots,0)\}
\subset} \def\nni{\supset} \def\und{\underline \fP_{\bf w}\times \fP_{\bf v},\] where we have used the abbreviations
$\fP_{\bf w}$ and $\fP_{\bf v}$ for the projective spaces above; similarly
we shall use the abreviation $\fP_{\bf w,v}$ for the image projective
space. That $\Phi$ is well-defined as claimed holds because the variables
$z_1,\ldots, z_n$ (resp. $t_1,\ldots, t_m$) have weights all divisible by
$v_0$ (resp. by $w_0$), hence a change of the branch of the $w_0^{th}$
roots of $y_0$ (resp. of the $v_0^{th}$ roots of $x_0$) just amounts to an
admissible overall scaling of the coordinates. The locus $\{y_0=x_0=0\}$
is the locus where all image values are zero, hence this consitutes the
locus where $\Phi$ is not defined (put differently, where $\Phi$ is not a
morphism but only a rational map). If we restrict $\Phi$ to $V_1\times
V_2$, then $x_0^{\ell} = -p(x_1,\ldots, x_n)$ and
$y_0^{\ell}=-q(y_1,\ldots, y_m)$. Hence
\begin{eqnarray*} p(z_1,\ldots, z_n)-q(t_1,\ldots, t_m) & = &
p(y_0^{w_1/w_0}\cdot x_1, \ldots, y_0^{w_n/w_0}\cdot x_n) -
q(x_0^{v_1/v_0}\cdot y_1, \ldots, x_0^{v_m/v_0}\cdot y_m) \\ & = &
y_0^{\nu/w_0}p(x_1,\ldots, x_n) - x_0^{\mu/v_0}q(y_1,\ldots, y_m) \\ & =
& y_0^{\ell}p(x_1,\ldots, x_n) -x_0^{\ell} q(y_1,\ldots, y_m) \\ & = &
-q(y_1,\ldots, y_m)\cdot p(x_1,\ldots, x_n) +p(x_1,\ldots, x_n)\cdot
q(y_1,\ldots, y_m) =0.
\end{eqnarray*}
Since $V_1\times V_2$ and $X$ both have dimension $n+m-2$, it is clear that
$\Phi_{|V_1\times V_2}$ is finite-to-one onto its image. \hfill $\Box$ \vskip0.25cm } \def\cD{\ifmmode {\cal D
We call the rational map $\Phi$ the {\it twist map}.
Let ${\bgm}_{\ell}$ denote the group of the $\ell$th roots of unity in
$\fC$.
\begin{corollary} Assume that gcd$(w_0,v_o,\ell)=1$. Then
$\Phi$ maps $V_1\times V_2$ to the quotient $V_1\times
V_2 / {\bgm}_{\ell}$, where ${\bgm}_{\ell}$ acts effectively
on $ V_1\times V_2 \subset} \def\nni{\supset} \def\und{\underline \fP_{(w_0,\bf
w)} \times \fP_{(v_0,\bf v)}$ by ($\gamma \in {\bgm}_{\ell}$)
\[ (\gamma,(x_0:\cdots :x_n),(y_0:\cdots:y_m)) \mapsto ((\gamma
x_0:x_1:\cdots:x_n), (\gamma y_0:y_1:\cdots:y_m)).\]
Hence the map $V_1\times V_2 \longrightarrow} \def\sura{\twoheadrightarrow X$ is generically $\ell : 1$.
\end{corollary}
{\bf Proof:} The product $V_1\times V_2$ is given by the two equations $\{
x_0^{\ell} + p(x_1,\ldots, x_n) = 0 = y_0^{\ell} + q(y_1,\ldots, y_m) \}$,
hence it is invariant under the given action. Suppose
$((x_0,\ldots,x_n),(y_0,\ldots, y_m))$ is a solution of the equations; then
for any $\gamma \in {\bgm}_{\ell}$, $((\gamma x_0,\ldots,x_n),(\gamma y_0,\ldots,
y_m))$ is also a solution. Provided the weights $w_0,\ v_0$ and $\ell$ have
no common divisor, none of the $\gamma$ act as an admissible overall scaling,
hence the result. \hfill $\Box$ \vskip0.25cm } \def\cD{\ifmmode {\cal D
\subsubsection{Resolving quotients}
It is well-known that the weighted projective spaces have only quotient
singularities (see \cite{dolg}, 1.2.5 and 1.3.3)
which can be resolved by the methods of torus
embeddings. Furthermore, from our assumption that $p$ and $q$ are
quasismooth, it follows that also $V$ has only quotient singularities (cf.\
\cite{dolg}, 3.1.6). This is then also true of the polynomial $p-q$
defining $X$, hence $X$ also has only quotient singularities, which can
again be resolved by torus methods.
We have seen that there is an action of ${\bgm}_{\ell}$ on $V$; we let
$\tilde{V}$ be a resolution of $V$ to which the action of ${\bgm}_{\ell}$
lifts. Let $\tilde{X}$ be a resolution of $X$. From these assumptions, the
map $\Phi$ lifts to a map $\tilde{\Phi}: \tilde{V} \longrightarrow} \def\sura{\twoheadrightarrow \tilde{X}$, and we have
the following commutative diagram
\[ \xymatrix{ \tilde{V}
\ar[r]^{\tilde{\Phi}} \ar[d] & \tilde{X}\ar[d] \\ V \ar@{
-->}[r]^{\Phi} & X.}\]
By assumption both $\tilde{V}$ and $\tilde{X}$ are
smooth.
\begin{corollary}\label{cquotient}
Under the above assumptions, $\~X$ is a resolution of the
quotients $V_1\times V_2/{\bgm}_{\ell}$ and $\tilde{V}/{\bgm}_{\ell}$.
\end{corollary}
\subsubsection{The fibration}
If we project the quotient $V_1\times V_2/{\bgm}_{\ell}$ onto the
individual factors, we get two {\it rational} fibrations, $\~X
\xymatrix{\ar@{-->}[r] & V_1/{\bgm}_{\ell}}
$ and $\~X \xymatrix{\ar@{-->}[r]& V_2/{\bgm}_{\ell} }$. In each case, the
generic smooth fibers are copies of resolutions of
$V_2$ (resp.\ $V_1$). We are especially
interested in the case that $\~X$ is Calabi-Yau.
\begin{lemma}\label{l3.4}
Suppose $\~X$ is Calabi-Yau and $w_0>1$. Then the rational
fibration $\~X \xymatrix{\ar@{-->}[r]& V_1/{\bgm}_{\ell}}$
induces a genuine fibration onto a resolution $Y$ of
$V_1/{\bgm}_{\ell}$, if and only if $\tilde{V}_2$ is also
Calabi-Yau.
\end{lemma}
{\bf Proof:} ``$\Ra$'': if $\~X \longrightarrow} \def\sura{\twoheadrightarrow Y$ is a fibration and $c_1(\~X) = 0$,
then by adjunction $c_1(F)= c_1(\~V_2) = c_1(\~X)_{|F} - c_1(N_{\~X}F) =0$.
``$\Leftarrow} \def\bs{\ifmmode {\setminus} \else$\bs$\fi$'': Suppose $c_1(\~X) =c_1(F)=c_1(\~V_2) =0$. Since by adjunction
this implies $c_1(N_{\~X}F)=0$, it suffices to show that $X \longrightarrow} \def\sura{\twoheadrightarrow
V_1/{\bgm}_{\ell}$ lifts to a morphism $\~X \longrightarrow} \def\sura{\twoheadrightarrow Y$ for a resolution $Y$
of $V_1/{\bgm}_{\ell}$. The map $X\longrightarrow} \def\sura{\twoheadrightarrow V_1/{\bgm}_{\ell}$ is given by
fixing $(x_1:\cdots:x_n)$ and mapping all $(y_0:\cdots :y_m)$ with
$\Phi(x,y) \in X$ to $(x_1:\cdots:x_n)$. Thus the projection is
well-defined unless $y_0=0$, a locus which is however blown up upon
resolution of $X$, provided $w_0>1$. \hfill $\Box$ \vskip0.25cm } \def\cD{\ifmmode {\cal D
\begin{corollary} \label{cfiber} Let $V_2$ and $X$ fulfill the necessary
conditions for being Calabi-Yau, $\sum_{j=0}^m v_j = \ell v_0$ and
$v_0\sum_{i=1}^n w_i + w_0 \sum_{j=1}^m v_j = v_0 w_0 \ell$. Then $X$ has
a resolution of singularities $\~X$ which is a fiber space over $Y$ with
constant modulus.
\end{corollary}
We now consider the possible singular fibers which can occur.
\begin{lemma}\label{l3.9}
The singular fibers of $\~X \longrightarrow} \def\sura{\twoheadrightarrow Y$ occur at the (image in
$V_1/{\bgm}_{\ell}$ of the) set of fixed points of the ${\bgm}_{\ell}$
action on $V_1$.
\end{lemma}
{\bf Proof:} This is well-known. \hfill $\Box$ \vskip0.25cm } \def\cD{\ifmmode {\cal D
We must consider two situations. First, if $x_0 =0$, then the entire locus
$\~{\Delta} = \{ p(x_1,\ldots,x_n)=0\} \subset} \def\nni{\supset} \def\und{\underline V_1$ is contained in the
discriminant and maps to $\Delta \subset} \def\nni{\supset} \def\und{\underline
V_1/{\bgm}_{\ell}$, and we let $\rho^*(\Delta)$ denote the total transform on
$Y$, where $\rho:Y\longrightarrow} \def\sura{\twoheadrightarrow V_1/{\bgm}_{\ell}$ is the resolution induced by that
of $\fP_{(\bf v,\bf w)}$. Secondly, it can happen that for $x_0\neq 0$ we
have further fixed points, a phenomenon which however only occurs in the
case of weighted projective spaces (we will show examples below). It is
easy to see that this can only occur along loci of the type $x_i=0$ for
some $i\in \{1,\ldots,n\}$.
\noindent{\bf Remark:} {\it Finding the fibers of the fibration}
In the physics literature these fibrations and the fibers are often found
by the method of ``eliminating coordinates''. As this is a source of
confusion, we remark here on this. Sometimes it works, but often one gets
misleading results. Consider the example of the K3 surface given
by the equation
\[ X= \{ z_1^{12}+z_2^6+z_3^4+z_4^2 =0 \}\subset} \def\nni{\supset} \def\und{\underline \fP_{(1,2,3,6)}.\]
According to our twist map, this is the quotient of the product
\[ \{x_0^4+x_1^{12}+x_2^6 =0\}\times \{ y_0^4+y_1^4+y_3^2=0\} \subset} \def\nni{\supset} \def\und{\underline
\fP_{(3,1,2)} \times \fP_{(1,1,2)}.\]
The latter curve is elliptic, and by our results above, the quotient of the
product by ${\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}/4{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}$ has a fibration with constant fibers equal to that
elliptic curve.
For $\hbox{\boldmath$\lambda$}}\def\bgm{\hbox{\boldmath$\mu$} = (\lambda_1:\lambda_2)\in \fP_{(1,2)}$ let $D_{\hbox{\scriptsize\boldmath$\lambda$}}=\{
\lambda_2z_1^2-\lambda_1^2 z_2 =0 \} \subset} \def\nni{\supset} \def\und{\underline \fP_{(1,2,3,6)}$. Let $X_{\hbox{\scriptsize\boldmath$\lambda$}} = X\cap
D_{\hbox{\scriptsize\boldmath$\lambda$}}$. This actually defines the fibration we already derived above,
over the weighted projective line $\fP_{(1,2)}$ (which is of course
isomorphic to the usual $\fP^1$). The equation for the fiber can be derived
by eliminating $z_2$, but upon eliminating $z_1$ (which occurs
quadratically in the equation of $D_{\hbox{\scriptsize\boldmath$\lambda$}}$), one gets the equation of a
{\it quotient} of the fiber. These equations are:
\[X_{\hbox{\scriptsize\boldmath$\lambda$}} =\{ (1+\left({\lambda_1^2\over
\lambda_2}\right)^6)z_2^6+z_3^4+z_4^2=0\}\subset} \def\nni{\supset} \def\und{\underline \fP_{(2,3,6)}\cong
\{(1+\left({\lambda_1^2\over \lambda_2}\right)^6)
(z_2')^2+(z_3')^2+z_4^2=0\}\subset} \def\nni{\supset} \def\und{\underline \fP_{(1,1,1)},\]
which describes a rational curve (a quotient of the actual fiber), and
\[X_{\hbox{\scriptsize\boldmath$\lambda$}}=\{((1+\left({\lambda_2\over
\lambda_1^2}\right)^6)z_1^{12}+z_3^4+z_4^2=0\}\subset} \def\nni{\supset} \def\und{\underline
\fP_{(1,3,6)}\cong\{(1+\left({\lambda_2\over
\lambda_1^2}\right)^6)
(z_1')^4+z_3^4+z_4^2=0\}\subset} \def\nni{\supset} \def\und{\underline \fP_{(1,1,2)},\]
which describes the elliptic curve which is the fiber. The rule to follow
is then: if one of the weights of the eliminated variables is one,
and the variable which is eliminated occurs {\it linearly} in the expression of
the divisor $D_{\hbox{\scriptsize\boldmath$\lambda$}}$, then one gets the correct answer.
To see that one must have one of the weights equal to unity, we consider
one more example. Consider the product of two weighted hypersurfaces of
degrees 6 and 12, respectively, in the product $\fP_{(2,1,1)}\times
\fP_{(4,1,1,6)}$, and let ${\bgm}_3$ act as above on this product
$V_1\times V_2$. Clearly $V_2$ is a K3 surface, and we get a fibration over
$\fP_{(1,1)}$, whose fibers are isomorphic to $V_2$. By the twist map this
maps to a hypersurface of degree 24 in $\fP_{(4,4,2,2,12)}$, which is
isomorphic to a hypersurface of degree 12 in
$\fP_{(2,2,1,1,6)}$. Note that in the equation of the
corresponding divisor $D_{\hbox{\scriptsize\boldmath$\lambda$}}$, both coordinates $z_1$ and $z_2$ occur
linearly, but both have weight 2. Upon elimination of either, we get a
hypersurface of degree 12 in $\fP_{(2,1,1,6)}$, which is of general type
and not K3. However, it is easy to see that it is a $2-1$ {\it cover} of
the K3 we started with. Hence the linearity of the coordinate in the
equation of the divisor is not sufficient.
\subsection{K3 surfaces}
We now apply the results above to the construction of K3 surfaces as
quotients of the type $C\times E$, where $C$ is a curve with an action of
${\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}/\ell{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}$, and $E$ is an elliptic curve with the same automorphism. To
apply the above, we must present both as weighted hypersurfaces.
\subsubsection{Elliptic curves} We consider the following elliptic curves.
\begin{eqnarray*} E_1 & = & \{ y_0^3 +y_1^3 +y_2^3 = 0 \} \subset} \def\nni{\supset} \def\und{\underline
\fP_{(1,1,1)}=\fP^2. \\ E_2 & = & \{y_0^4+y_1^4+y_2^2 = 0 \} \subset} \def\nni{\supset} \def\und{\underline
\fP_{(1,1,2)}. \\ E_3 & = & \{y_0^6+y_1^3+y_2^2=0\} \subset} \def\nni{\supset} \def\und{\underline \fP_{(1,2,3)}.
\end{eqnarray*}
Both $E_1$ and $E_3$ are elliptic curves with modulus $\tau=\varrho=e^{2\pi
i/3}$, while $E_2$ has $\tau = i$.
\subsubsection{K3 surfaces}\label{sK3}
We let $C_{(w_0,w_1,w_2)}$ denote the
following curve
\begin{equation}\label{11} C_{(w_0,w_1,w_2)} = \{x_0^{\ell} +
p(x_1,x_2)=0\} \subset} \def\nni{\supset} \def\und{\underline \fP_{(w_0,w_1,w_2)}
\end{equation}
of degree $\ell\cdot w_0$. Then applying the map $\Phi$ of Lemma \ref{l3.1}
we get a rational map of $C_{(w_0,w_1,w_2)} \times E_i$ (where $i=1$ or 3
if $\ell=3$ or 6, respectively,
and $i=2$ for $\ell =4$) onto a hypersurface $X\subset} \def\nni{\supset} \def\und{\underline
\fP_{(v_0w_1,v_0w_2,w_0v_1,w_0v_2)}$ of degree $d$, where $(v_0,v_1,v_2)=
(1,1,1),\ (1,1,2)$ and $(1,2,3)$ for the three elliptic curves. We list
some of the possible hypersurfaces, all of which are of Fermat type,
that one gets in this manner. Assuming that
$d=\sum k_i$, where $(k_1,k_2,k_3,k_4)=(v_0w_1,v_0w_2,w_0v_1,w_0v_2)$, we
necessarily have that $X$ is a (singular) K3 surface, and $\tilde{X}$ is
its minimal desingularisation. The singular fibers we list are those
occuring if the polynomial $p$ in the definition of the curve
$C_{(w_0,w_1,w_2)}$ is of the form $p(x_1,x_2)=x_1^{\ell\cdot w_0/w_1}+
x_2^{\ell\cdot w_0/w_2}$. In the last case the polynomials are not of
Fermat type.
\[ \begin{array}{|c|c|c|c|c|c|c|} \hline
\# & (w_0,w_1,w_2) & (v_0,v_1,v_2) & \ell &
(k_1,k_2,k_3,k_4) & d & \hbox{singular fibers} \\ \hline \hline
1 & (2,1,1) & (1,1,1) & 3 & (1,1,2,2) & 6 & 6\times IV \\ \hline
2& & (1,1,2) & 4 & (1,1,2,4) & 8 & 8\times III \\ \hline
3& & (1,2,3) & 6 & (1,1,4,6) & 12 & 12\times II \\ \hline
4& (3,1,2) & (1,1,2) & 4 & (1,2,3,6) & 12 & 6\times III, 1\times I_0^* \\ \hline
5& & (1,2,3) & 6 & (1,2,6,9) & 18 & 9\times II, 1\times I_0^* \\ \hline
6& (4,1,3) & (1,1,1) & 3 & (1,3,4,4) & 12 & 4\times IV, 1\times IV^* \\ \hline
7& & (1,2,3) & 6 & (1,3,8,12)& 24 & 8\times II, 1\times IV^* \\ \hline
8& (5,1,4) & (1,1,2) & 4 & (1,4,5,10)& 20 & 5\times III, 1\times III^* \\ \hline
9& (7,1,6) & (1,2,3) & 6 & (1,6,14,21)&42 & 7\times II, 1\times II^* \\ \hline
10& (5,2,3) & (1,2,3) & 6 & (2,3,10,15)&30 & 5\times II, 1\times IV^*,
1\times I_0^*\\ \hline
11& (11,5,6) & (1,2,3) & 6 & (5,6,22,33) & 66 & 2\times II,\ 2\times II^*
\\ \hline
\end{array}\]
In a sequel to this paper we will describe in detail how one gets the
singular fibers listed. Let us just make a few remarks about the example
11. For the weights in this case, a Fermat hypersurface is
not possible. We consider instead the following polynomial:
\[
\{z_0^{12}z_1+z_1^{11}+z_2^3+z_3^2=0\} \subset} \def\nni{\supset} \def\und{\underline \fP_{(5,6,22,33)}.\]
We see without difficulty that this is the image under the twist map
\begin{eqnarray*} \fP_{(11,5,6)}\times \fP_{(1,2,3)} & \longrightarrow} \def\sura{\twoheadrightarrow &
\fP_{(5,6,22,33)}\\((x_0:x_1:x_2),(y_0:y_1:y_2)) & \mapsto &
(y_0^{5/11}x_1: y_0^{6/11}x_2:x_0^2y_1:x_0^3y_2 )
\end{eqnarray*}
of the product $ \{ x_0^6+x_1^{12}x_2 +x_2^{11}
=0\} \times \{y_0^6+y_1^3+y_2^2=0\}$.
First we note that since $z_1$ only occurs in the
mixed monomial to first power, that the polynomial $p(z_0:z_1:z_2:z_3):=
z_0^{12}z_1+z_1^{11}+z_2^3+z_3^2$ is transversal. It follows that, since
the degree of $X:=\{p(z)=0\}$ is 66, which is also equal to the sum of the
weights, that $X$ is a K3 surface. Its proper transform, after resolving
the singularities, is a smooth K3 surface, which we for convenience also
denote by $X$. From the fact that it is the
image of the twist map above, it follows immediately that
$X$ has a constant-modulus elliptic fibration.
\subsubsection{An exotic surface}\label{exoticsurface}
We already mentioned that the condition
$d=\sum k_i$ is sufficient for $X$ to be K3. It is, however, not
necessary, which we show by example. We are indepted to Igor Dolgachev for
explaining this example to us.
Consider the weighted hypersurface
\[
x^2+y^3 +z^{11} + w^{66} =0\]
in ${\bf P}_{(1,6,22,33)}$. Here $d=66$ while $\sum k_i=62$, hence the
sufficient condition above is not met, and the hypersurface does not look
like a K3. By general principles of weighted hypersurfaces, the geometric
genus is given by the number of
monomials of degree $66-62$, of which there is only
one, namely, $w^4$. It follows that a canonical divisor is given by the
locus $w=0$, which is a curve isomorphic to
\begin{equation}\label{12} C=\{z^{11}+y^3+x^2 = 0\} \subset} \def\nni{\supset} \def\und{\underline \fP_{(6,22,33)}.
\end{equation}
The resolution of such a weighted hypersurface is well-known, see
\cite{OW}. One considers
continued fractions of rational numbers derived from the equation of the
hypersurface as follows. Let $(k_1,k_2,k_3,k_4)$ be the weights and for
each $i=1,\ldots,4$, $m_i$ the corresponding degrees of the monomial
$x_i^{m_i}$ occuring in the equation of the hypersurface. Then, to each
singular point, with local stabilizer ${\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}_{\alpha_i}$, the rational number
of which one considers the continued fraction is ${\alpha_i\over \beta_i}$,
where $\beta_i$ is a solution of $k_i\cdot \beta_i = 1mod(\alpha_i)$. In
our case, the $\alpha_i$'s are just the $m_i$'s, and the three fractions we
need to expand are ${2\over 1},\ {3\over 1},\ {11\over 2}$. The first two
give rise to a single $-2$ and $-3$ curve, respectively, while the latter
gives a $-6$ curve meeting a $-2$ curve. Together with the curve $C$ given
here by $x^2+y^3+z^{11}=0$ we have the following resolution of the three
singular points on the K3 surface:
\[ \epsfbox{reso.eps} \]
The self-intersection number of the central curve $C$ (the proper transform
of the hyperplane section $w=0$ above) can be determined as follows. Letting
$(\alpha_i,\beta_i)$ denote the pairs giving the continued fractions as
above, the formula is (cf \cite{OW}, 3.6.1, which do the case of an isolated
singularity, but the proof is the same; note that the formula in {\it
loc.~cit.} should have a minus sign in front of the first term):
the self-intersection number is
$-b$, where $b$ is given by
\[ b=-{1\over k_2k_3k_4} + \sum {\beta_i\over \alpha_i}.\]
This yields in our case
\[ b = -{1\over 66} + {1\over 2} + {1\over 3} + {2\over 11} =1.\]
Hence, the central curve $C$ is {\it exceptional of the first kind} and can
be blown down. The $-2$ curve becomes a $-1$ curve and can be blown down,
then the $-3$ curve of the original curve is now a $-1$ curve and can be
blown down. The result is depicted below.
\[ \epsfbox{birat.eps} \]
This leaves two curves $F$ and $D$, where
$F$ is a cuspidal genus 1 curve and $D$ is a $-2$ curve intersecting it
transversally. The linear system $|F|$ is an elliptic pencil and $D$ its
section.
Letting $S$ denote this smooth surface, it is clearly an elliptic K3
surface. It admits an action of ${\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}/66{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}$ which lifts the action
$(x,y,z,\zeta_{66}w)$ on the weighted hypersurface.
We can describe this fibration as follows. Consider the subspace $\{z=w=0\}
\cong \fP_{(22,33)}\cong \fP^1
\subset} \def\nni{\supset} \def\und{\underline \fP_{(1,6,22,33)}$. The set of hypersurfaces of
minimal degree containing this $\fP^1$ are given by the equations $D_{\lambda}
= \{ z-\lambda w^6 =0\}$ (for $\lambda\neq 0$). The intersection with the
hypersurface $X$ is then
\begin{eqnarray*}
X_{\lambda} := D_{\lambda}\cap X & = & \{ x^2+y^3+(1+\lambda^{11})w^{66} =0\} \cong
\fP_{(1,22,33)}[66] \\ & \cong & \{ x^2+y^3+(1+\lambda^{11})(w')^6 =0\} =
\fP_{(1,2,3)}[6],
\end{eqnarray*}
which is the elliptic curve of degree 6 in $\fP_{(1,2,3)}$, which is
isomorphic to the Fermat curve, which, as we pointed out above, is the
elliptic curve $E_{\rho}$ with modulus $\rho = e^{2\pi i \over 3}$. The
proper transforms of the
$X_{\lambda}$ are the fibers of our fibration. We need also
the action of ${\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}/66{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}$ on the holomorphic two-form on the K3
surface. The action on the fibers is:
\[ w \mapsto \zeta_{66} w \Ra w' \mapsto \zeta_{66}^{11} w' =
\zeta_6 w',\]
which is the usual action of ${\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}/6{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}$ on the sextic in
$\fP_{(1,2,3)}$. The action on the base of the fibration is the action on
the parameter $\lambda$ (on some Zariski open set) above,
\[ \lambda w^6 \mapsto \zeta_{11} \lambda w^6,\ i.e.,\ z\mapsto \zeta_{11} z,\
\lambda\mapsto \zeta_{11} \lambda,\]
and the action on the base is just multiplication by $\zeta_{11}$, which
permutes the 11 singular fibers at the 11th roots of $-1$ and fixes the
singular fiber at $\infty$. Since in terms of local coordinates $(z,t)$,
where $z$ is a fiber coordinate, $t$ a coordinate on the base, the
holomorphic two-form is given by $dz\wedge dt$, it follows that
\[ dz \wedge dt = dw \wedge d\lambda \mapsto \zeta_6 dw \wedge \zeta_{11} d\lambda
= \zeta_{66} dz\wedge dt,\]
and the action is by multiplication by $\zeta_{66}$, i.e., if $h$ denotes
the generator of ${\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}/66{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}$ and $\omega} \def\cO{{\cal O}} \def\gG{\Gamma} \def\fQ{{\bf Q}$ denotes the holomorphic two-form,
then
\begin{equation}\label{e66} h^*(\omega} \def\cO{{\cal O}} \def\gG{\Gamma} \def\fQ{{\bf Q}) = \zeta_{66} \omega} \def\cO{{\cal O}} \def\gG{\Gamma} \def\fQ{{\bf Q}.
\end{equation}
This surface is known
from the work of Kondo. He shows in \cite{Kondo2} that there is in fact a
{\it unique} K3 surface which admits ${\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}/66{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}$ as an automorphism
group which preserves the Picard group of the K3 surface, or equivalently,
acts non-trivially on the holomorphic two-form (see section 3.3.2 below).
Kondo describes the same surface as an elliptic surface
with 12 fibers of type \II\
at $t=0$ and at the $11^{th}$
roots of unity. The affine equation is
\[y^2=x^3+t\prod_1^{11}(t-\zeta_{11}^i),\]
and the automorphism is given by
\[ g: (x,y,t) \mapsto (\zeta_{66}^2 x, \zeta_{66}^3 y, \zeta_{66}^6 t).\]
He shows that the Picard lattice of this surface is just a hyperbolic
plane, while the transcendental lattice is of the form $U\oplus U \oplus E_8
\oplus E_8$, where $E_8$ denotes the negative definite copy of the root
lattice of the exceptional group of type $E_8$. For the elliptic surface in
Weierstra\ss\ form, the holomorphic two-form is given by ${dx\over y}
\wedge dt$, and it follows that $g^*(\omega} \def\cO{{\cal O}} \def\gG{\Gamma} \def\fQ{{\bf Q})=\zeta_{66}^5 \omega} \def\cO{{\cal O}} \def\gG{\Gamma} \def\fQ{{\bf Q}$, which is a
primitive $66$th root of unity, showing that the action of the group of
order 66 acts trivially on the Picard group. Comparing with (\ref{e66}), we
see that $h=g^{55}$.
\subsection{Calabi-Yau threefolds}
In this section we apply the twist map to construct Calabi-Yau threefolds
with K3-fibrations, which are quotients of either products $S\times E$,
where $S$ is a surface and $E$ is an elliptic curve, or products
$C\times K$, where $K$
is now a K3 surface. To apply the latter, we need some information on the
automorphisms of K3 surfaces. This information can be found in
\cite{Kondo2} and \cite{Kondo1}. We first recall this, then pass to the
construction.
\subsubsection{Elliptic fibrations}\label{selliptic}
We do not attempt to give a
classification, but instead describe a few examples. Lists of such elliptic
weighted hypersurface Calabi-Yau threefolds have been compiled, and for
many of these we can realize them as quotients. As before we use the three
elliptic curves of section 3.2.1 but instead of the curve (\ref{11}) now
the variety $V_1$ is the surface
\begin{table}\[ \begin{array} {|l|l| c | l | c|} \hline (w_0,w_1,w_2,w_3,)
& (v_0,v_1,v_2) & \ell & (k_1,k_2,k_3,k_4,k_5) & d \\ \hline \hline
(3,1,1,1) & (1,1,1) & 3 & (1,1,1,3,3) & 9 \\
& (1,1,2) & 4 & (1,1,1,3,6) & 12 \\
& (1,2,3) & 6 & (1,1,1,6,9) & 18 \\ \hline
(4,1,1,2) & (1,1,1) & 3 & ( 1,1,2,4,4) & 12 \\
& (1,1,2) & 4 & (1,1,2,4,8) & 16 \\
& (1,2,3) & 6 & (1,1,2,8,12) & 24 \\ \hline
(5,1,1,3) & (1,1,1) & 3 & (1,1,3,5,5) & 15 \\
& (1,2,3) & 6 & (1,1,3,10,15) & 30 \\ \hline
(5,1,1,2) & (1,1,2) & 4 & (1,2,2,5,10) & 20 \\
& (1,2,3) & 6 & (1.2.2.10.15) & 30 \\ \hline
(6,1,1,4) & (1,1,2) & 4 & (1,1,4,6,12) & 24 \\
& (1,2,3) & 6 & (1,1,4,12,18) & 36 \\ \hline
(6,1,2,3) & (1,1,1) & 3 & (1,2,3,6,6) & 18 \\
& (1,1,2) & 4 & (1,2,3,6,12) & 24 \\
& (1,2,3) & 6 & (1,2,3,12,18) & 36 \\ \hline
(7,1,2,4) & (1,1,2) & 4 & (1,2,4,7,14) & 28 \\ \hline
(7,1,3,3) & (1,1,1) & 3 & (1,3,3,7,7) & 21 \\
& (1,2,3) & 6 & (1,3,3,14,21) & 42 \\ \hline
(7,2,2,3) & (1,2,3) & 6 & (2,2,3,14,21) & 42 \\ \hline
(8,1,1,6) & (1,1,1) & 3 & (1,1,6,8,8) & 24 \\
& (1,2,3) & 6 & (1,1,6,16,24) & 48 \\ \hline
(8,1,3,4) & (1,2,3) & 6 & (1,3,4,16,24) & 48 \\ \hline
(8,2,3,3) & (1,2,3) & 6 & (2,3,3,16,24) & 48 \\ \hline
(9,1,2,6) & (1,1,2) & 4 & (1,2,6,9,18) & 36 \\
& (1,2,3) & 6 & (1,2,6,18,27) & 54 \\ \hline
(9,1,4,4) & (1,1,2) & 4 & (1,4,4,9,18) & 36 \\ \hline
(9,2,3,4) & (1,1,2) & 4 & (2,3,4,9,18) & 36 \\ \hline
(10,1,1,8)& (1,1,2) & 4 & (1,1,8,10,20) & 40 \\ \hline
(10,2,3,5) & (1,1,1) & 3 & (2,3,5,10,10) & 30 \\
& (1,2,3) & 6 & (2,3,5,20,30) & 60 \\ \hline
(10,1,3,6) & (1,1,1) & 3 & (1,3,6,10,10) & 30 \\
& (1,2,3) & 6 & (1,3,6,20,30) & 60 \\ \hline
(10,3,3,4) & (1,2,3) & 6 & (3,3,4,20,30) & 60 \\ \hline
(12,1,2,9) & (1,1,1) & 3 & (1,2,9,12,12) & 36 \\
& (1,2,3) & 6 & (1,2,9,24,36) & 72 \\ \hline
(13,1,6,6) & (1,2,3) & 6 & (1,6,6,26,39) & 78 \\ \hline
(14,1,1,12) & (1,2,3) & 6 & (1,1,12,28,42) & 84 \\ \hline
\end{array}\]\caption{List of elliptic Calabi-Yau threefold
weighted hypersurfaces of Fermat type and with constant fiber modulus}
\end{table}
\begin{equation}\label{surface} S_{(w_0,w_1,w_2,w_3)} =\{ x_0^{\ell} +
p(x_1,x_2,x_3) = 0 \}
\end{equation}
of degree $\ell w_0$. Applying the map $\Phi$ of Lemma \ref{l3.1} we get a
rational map of $S_{(w_0,\ldots, w_3)} \times E_i$ (where $i=1$ or 3
if $\ell=3$ or 6, respectively, and $i=2$ for $\ell =4$) onto a threefold of
degree $d=\ell w_0$. again we list some of the possiblities which a brief
manual search comes up with. All these cases have the following properties:
they are of Fermat type (quite restrictive) and they have a fibration of
{\it constant modulus} (it is not clear how restrictive this is).
This list is given in Table 1.
As an explicit example consider the case $(2k,1,1,2(k-1)),\ (1,2,3)$ for
which the Calabi-Yau is the hypersurface of degree $12k$ in
$\fP_{(1,1,2(k-1),4k,6k)}$. The elliptic fibration is onto
$\fP_{(1,1,2(k-1))}$, whose desingularisation is the Hirzebruch surface
$\fF_{2(k-1)}$. By Lemma \ref{l3.9}, the discriminant is the total
transform of $\Delta=\{p(x_1,x_2,x_3)=0\} \subset} \def\nni{\supset} \def\und{\underline \fP_{(1,1,2(k-1))}$. The
projection $\fP_{(1,1,2(k-1))} \longrightarrow} \def\sura{\twoheadrightarrow \fP_{(1,1)}$ lifts to the projection
$\fF_{2(k-1)} \longrightarrow} \def\sura{\twoheadrightarrow \fP^1$. Let $C_{\infty}$ be the section of negative
self-intersection $-2(k-1)$ (the exceptional curve of the resolution
$\fF_{2(k-1)} \longrightarrow} \def\sura{\twoheadrightarrow \fP_{(1,1,2(k-1))}$), and let $C_0$ denote the class of
a positive section, $F$ the class of a fiber. The curve $\Delta$ is reducible;
its total transform on $\fF_{2(k-1)}$ consists of an irreducible curve $C$
of class $C\sim aC_0+bF$ ($a,b>0$) and a multiple of $C_{\infty}$, $\Delta
\sim a C_0 + bF + \nu C_{\infty}$. Assume that the exceptional point
$(0,0,1)$ of $\fP_{(1,1,2(k-1))}$ is not contained in $\Delta$ (for example
$p$ of Fermat type). Then $C$ and $C_{\infty}$ are disjoint. Hence given
the explicit form of $p$ one can easily determine exactly the degeneracy
locus of the smooth elliptic Calabi-Yau.
In the cases $k=2,3,4,7$ and $\ell=6$ or $k=5$ and $\ell =4$
one can take $p$ to be of Fermat type. In these
cases, the class of $C$ can be calculated as follows: it is a curve in
$\fP_{(1,1,2(k-1))}$ of degree $12k$,
which maps $k$ to one onto a $\fP^1$, totally branched at $12k$ points,
hence the Euler number (which is the first Chern number) is equal to
$k\cdot(2-2k)+2k=2k(2-k)$. Then applying adjunction on the resolution
$\fF_{2(k-1)}$, one gets $a=k $ and $b=-2k(k-1) $. The fiber over
$C_{\infty}$ is of type IV, I$_0^*$, IV$^*$ and II$^*$ in the $\ell=6$
cases and III$^*$ in the $k=5,\ \ell=4$ case, and this determines $\nu$ to
be 3, 6, 8, 10 and 9 in the respective cases. These have been studied in
more detail in \cite{MV}.
\begin{lemma}\label{l5} Let $X$ be a Calabi-Yau threefold with both an
elliptic and a K3 fibration, and suppose these are compatible (i.e., that
the elliptic fibration of $X$ restricts to an elliptic fibration of the
smooth fibers), and assume moreover that the degenerate fibers of the K3
fibration are irreducible. Then the base of the elliptic fibration is a
rational ruled surface $\bf F_n$.
\end{lemma}
{\bf Proof:} From the first assumption, we clearly have on the base of the
elliptic fibration, for each K3 fiber, a $\fP^1$, the base of the elliptic
fibration of that K3 fiber. Thus the surface fibers over the original
$\fP^1$, and is thus the blow up of one of the Hirzebruch surfaces. Now we
use the assumption that each of the degenerate fibers is irreducible; being
so, each of these can fiber over at most a single $\fP^1$, and it follows
that {\it all} fibers of the base surface are just $\fP^1$, in other words
that it is a rational ruled surface, as claimed. \hfill $\Box$ \vskip0.25cm } \def\cD{\ifmmode {\cal D
In general the base of such a fibration can be some blow-up of one of the
$\fF_n$.
\begin{table}
\[ \begin{array}{|l|l||c|c|c|} \hline
(w_0,w_1,w_2,w_3) & (k_1,k_2,k_3,k_4,k_5) & d & \chi &
h^{1,1} \\ \hline \hline
(581,41,42,498) & (41,42,498,1162,1743) & 3486 & 960 & 491 \\ \hline
(498,36,41,421) & (36,41,421,996,1494) & 2988 & 960 & 491 \\ \hline
(539,36,41,462) & (36,41,462,1078,1617) & 3234 & 900 & 462 \\ \hline
(469,31,42,396) & (31,42,396,938,1407) & 2814 & 900 & 462 \\ \hline
(463,31,41,391) & (31,41,391,926,1389) & 2778 & 900 & 462 \\ \hline
(433,31,36,366) & (31,36,366,866,1299) & 2598 & 840 & 433 \\ \hline
(483,28,41,414) & (28,41,414,966,1449) & 2898 & 804 & 416 \\ \hline
(414,24,41,349) & (24,41,349,828,1242) & 2484 & 804 & 416 \\ \hline
(385,28,31,326) & (28,31,326,770,1155) & 2310 & 744 & 387 \\ \hline
(434,21,41,372) & (21,41,372,868,1302) & 2604 & 720 & 377 \\ \hline
(372,18,41,313) & (18,41,313,744,1116) & 2232 & 720 & 377 \\ \hline
\end{array}\]
\caption{\label{table2}
Cases of elliptic fibrations with constant modulus (fiber the
sextic elliptic curve in $\fP_{(1,2,3)}$), for which the Euler number
$\chi$ is large, positive}
\end{table}
We now turn to some examples of elliptic fibrations which are not K3
fibrations. The examples given up to now were all in the geography region
where the Euler number is negative. We now give some examples where the
Euler number is positive, which are listed in Table \ref{table2}.
We will discuss one example in more detail.
These examples are all images of the twist map
\[ \fP_{(w_0,w_1,w_2,w_3)} \times \fP_{(1,2,3)} \longrightarrow} \def\sura{\twoheadrightarrow
\fP_{(w_1,w_2,w_3,2\cdot w_0,3\cdot w_0)}.\]
Hence they all have constant elliptic fibrations with fiber the sextic in
$\fP_{(1,2,3)}$. We now consider the first example in some more detail.
The variety $V_1$ is a weighted hypersurface in $\fP_{(581,41,42,498)}$ of
degree 3486, given by the equation
$$x_0^6+p(x_1,x_2,x_3) = x_0^{6}+x_1^{84}x_2 + x_2^{83} +
x_3^{7} = 0$$ and is a $6-1$ cover of the weighted $\fP_{(41,42,498)} \cong
\fP_{(41,7,83)}$, branched over the locus $p=0$. Then the base of the
elliptic fibration will be the resolution of $\fP_{(41,7,83)}$. This is
standard toric geometry. Take the lattice spanned by the three vectors
\[ v_0 = {1\over 41} \left( \begin{array}{c} -1 \\ -1
\end{array} \right), \quad v_1={ 1\over 7} \left(\begin{array}{c} 1 \\ 0
\end{array} \right), \quad v_2={1\over 83}\left(\begin{array}{c} 0 \\ 1
\end{array} \right), \]
and in each of the cones spanned by two of the $v_i$ one must determine all
lattice points. Then taking the convex hull of these, one gets a simplicial
decomposition of the cones into subcones all of which have unit area. Then
one counts the number of new vertices, and each corresponds to an
exceptional divisor. There are three singular points, and our curve $p=0$
passes through only $(1,0,0)$, so this is the only relevant singularity.
It is easy to see that we are looking for integral solutions
$(\alpha,\beta,\gamma)$ of the
inequalities
\[ \alpha(w_1+w_2)-w_0(\beta+\gamma) + w_0 \leq 0,\quad w_0\beta>w_1\alpha,\quad
w_0\gamma>w_2\alpha,\]
where $(w_0,w_1,w_2)=(41,7,83)$,
of which there are the following 20,
in other words, there are 20 exceptional curves
resolving the point:
\begin{eqnarray*} (\alpha,\beta,\gamma) & = & (11, 2, 23), (16, 3, 33), (17, 3,
35), (21, 4, 43), (22, 4, 45), (23, 4, 47), (26, 5, 53), \\ & &
(27, 5, 55), (28, 5, 57), (29, 5, 59), (31, 6, 63), (32, 6, 65, )(33, 6,
67), (34, 6, 69), \\ & &
(35, 6, 71), (36, 7, 73), (37, 7, 75), (38, 7, 77), (39, 7, 79), (40, 7, 81).
\end{eqnarray*}
It follows that along the proper transform of $p=0$,
we have singularities of type II, while along the 20 exceptional curves we
have singularities of type II$^*$. The morphism onto the base is not yet
flat; to achieve this one could take Miranda's small resolutions, which
would amount to blowing up the intersection points of the 20 exceptional
curves.
\subsubsection{Automorphisms of K3 surfaces}
This section is prepatory and recalls some facts about the automorphism of
K3 surfaces, which will be applied in the same way the automorphisms of
orders 2, 3, 4 and 6 were for elliptic curves.
Let $X$ be a K3 surface; the integral homology ${\rm H}^2(X,{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P})$ forms a lattice
in ${\rm H}^2(X,\fR)\cong \fR^{22}$ which is a copy of $U\oplus U \oplus U \oplus
E_8 \oplus E_8$, where $E_8$ denotes here the negative definite even
unimodular lattice in $\fR^8$ and $U$ is the hyperbolic lattice. We
consider in this note {\it polarized} K3 surfaces, letting $\omega \in
{\rm H}^{1,1}(X)$ denote the K\"ahler form; the primitive cohomology
${\rm H}^2(X,\fC)_0$ is then the orthocomplement in ${\rm H}^2(X,\fC)$ to
$<\omega>$. By an {\it automorphism} of $X$ we will mean in this paper a
polarization-preserving automorphism $\sigma:X\longrightarrow} \def\sura{\twoheadrightarrow X$; we let also
$\sigma_*$ denote the induced map on cohomology $\sigma_*\in
\hbox{Aut}} \def\Im{\hbox{Im}} \def\mod{\hbox{mod}({\rm H}^2(X,{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}))$. The map
\begin{eqnarray*} *: \hbox{Aut}} \def\Im{\hbox{Im}} \def\mod{\hbox{mod}(X) & \longrightarrow} \def\sura{\twoheadrightarrow & \hbox{Aut}} \def\Im{\hbox{Im}} \def\mod{\hbox{mod}({\rm H}^2(X,{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P})) \\ \sigma &
\mapsto & \sigma_*
\end{eqnarray*}
is injective (cf. \cite{Kondo1}, section 4) so it is convenient to consider
this as a representation of $\hbox{Aut}} \def\Im{\hbox{Im}} \def\mod{\hbox{mod}(X)$.
Let $S_X:={\rm H}^{1,1}(X) \cap {\rm H}^2(X,{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P})$ be the {\it Picard lattice} of $X$,
and let $T_X$ be its orthocomplement in ${\rm H}^2(X,{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P})$; this is the {\it
transcendental lattice} of $X$. The direct sum $S_X\oplus T_X$ is of
finite index in ${\rm H}^2(X,{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P})$, and the map $*$ actually maps into
$\hbox{Aut}} \def\Im{\hbox{Im}} \def\mod{\hbox{mod}(S_X\oplus T_X) \cong O(S_X)\times O(T_X)$ (cf.\ \cite{Kondo2},
Proposition 1.1), so we can faithfully consider the action of $\hbox{Aut}} \def\Im{\hbox{Im}} \def\mod{\hbox{mod}(X)$ on
the Picard and transcendental lattices. Note that if $T_X$ (and hence
$S_X$) is {\it unimodular}, then
\[ S_X\oplus T_X={\rm H}^2(X,{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P})=U^{\oplus 3}\oplus E_8^{\oplus 2},\]
so both of $S_X$ and $T_X$ are themselves direct sums with summands $U$ and
$E_8$.
Every element $g\in \hbox{Aut}} \def\Im{\hbox{Im}} \def\mod{\hbox{mod}(X)$ acts on this
decomposition, and letting $\alpha_g\in \fC^*$ denote the factor
such that $g^*(\Omega)=\alpha_g \Omega$, where $\Omega$ denotes the
holomorphic two form, the map $g\mapsto \alpha_g$
gives rise to an exact sequence
\[ 1 \longrightarrow} \def\sura{\twoheadrightarrow G_X \longrightarrow} \def\sura{\twoheadrightarrow \hbox{Aut}} \def\Im{\hbox{Im}} \def\mod{\hbox{mod}(X) \stackrel{\alpha}{\longrightarrow} \def\sura{\twoheadrightarrow} H_X \longrightarrow} \def\sura{\twoheadrightarrow 1.\]
Elements of $G_X$ are called {\it symplectic}, as they preserve the
symplectic form $\Omega$. If $X$ is algebraic, then $H_X={\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}_k$ for some
$k$. The following facts were proved by Nikulin in \cite{vvn}:
\begin{itemize}\item[1.] $G_X=\Ker(\hbox{Aut}} \def\Im{\hbox{Im}} \def\mod{\hbox{mod}(X) \longrightarrow} \def\sura{\twoheadrightarrow \hbox{Aut}} \def\Im{\hbox{Im}} \def\mod{\hbox{mod}(T_X))$ and
$H_X=\Ker(\hbox{Aut}} \def\Im{\hbox{Im}} \def\mod{\hbox{mod}(X)\longrightarrow} \def\sura{\twoheadrightarrow \hbox{Aut}} \def\Im{\hbox{Im}} \def\mod{\hbox{mod}(S_X))$. \item[2.] The representation
${\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}_k\longrightarrow} \def\sura{\twoheadrightarrow \hbox{Aut}} \def\Im{\hbox{Im}} \def\mod{\hbox{mod}(T_X)$ is a direct sum of a number
$K$ of irreducible representations,
each of rank $\phi(k)$ ($\phi$ denotes the Euler phi-function),
so $\sum^K_1\phi(k)=K \phi(k) =rank(T_X)$. \item[3.]
The representation $G_X\longrightarrow} \def\sura{\twoheadrightarrow \hbox{Aut}} \def\Im{\hbox{Im}} \def\mod{\hbox{mod}(S_X)$ was described by Nikulin for
abelian $G_X$; the possible subgroups occuring are
$$({\bf Z}_2)^m, \ 0\leq m\leq 4;\ \ {\bf Z}_4;\ \ {\bf Z}_2\times
{\bf Z}_4;\ \ ({\bf Z}_4)^2;\ \ {\bf Z}_8;\ \ {\bf Z}_3;$$
$$ ({\bf
Z}_3)^2;\ \ {\bf Z}_5;\ \ {\bf Z}_7; \ \ {\bf Z}_6;\ \ {\bf Z_2}\times
{\bf Z}_6.$$
\end{itemize}
From the fact that $\phi(k)\leq 22$ one deduces $k\leq 66$, and $k=66$ does
occur. This was considered by Kondo in \cite{Kondo2} and by Vorontsov in
\cite{Vor}, who proved the following results:
\begin{itemize}\item[1.] If $T_X$ is unimodular, then $k$ is a divisor of
66, 44, 42, 36, 28 or 12. \item[2.] Suppose $\phi(k)=rank(T_X)$.
Then $k$=66, 44, 42, 36, 28 or 12,
and in these cases there exists a unique (up to isomorphism) K3 surface
with given $k$. \item[3.] If $T_X$ is not unimodular, then $k=2^r\ (1\leq
r \leq 4),\ 3^r\ (1\leq r \leq 3),\ 5^r\ (r=1, 2),\ 7, 11, 13, 17$ or 19.
\end{itemize}
In the case that $T_X$ is unimodular, it is a direct sum of factors $U$ and
$E_8$. For the cases above, the lattices are listed in the following
table.
\[\begin{array}{c | c c} k & S_X & T_X \\ \hline
66 & U & U\oplus U \oplus E_8 \oplus E_8 \\
44 & U & U\oplus U \oplus E_8 \oplus E_8 \\
42 & U\oplus E_8 & U\oplus U \oplus E_8 \\
36 & U\oplus E_8 & U\oplus U \oplus E_8 \\
28 & U\oplus E_8 & U\oplus U \oplus E_8 \\
12 & U\oplus E_8 \oplus E_8 & U \oplus U
\end{array}
\]
\subsubsection{Calabi-Yau threefolds with K3 fibrations}
Again we just list some examples. As above we consider three K3's of the
above list.
\begin{eqnarray*} K_1 & = & \{y_0^6+y_1^6+y_2^3+y_3^3 =0\}\subset} \def\nni{\supset} \def\und{\underline
\fP_{(1,1,2,2)} \\
K_2 & = & \{y_0^{12}+y_1^6+y_2^4+y_3^2=0\} \subset} \def\nni{\supset} \def\und{\underline
\fP_{(1,2,3,6)} \\
K_3 & = & \{y_0^{42}+y_1^7+y_2^3+y_3^2 =0 \} \subset} \def\nni{\supset} \def\und{\underline
\fP_{(1,6,14,21)}.
\end{eqnarray*}
All three K3's are elliptic fibrations;
the elliptic fibers are $E_1,\ E_2$ and $E_3$, respectively, and $\ell = 6,
12, 42$. We again consider the curve (\ref{11}), and the weighted
hypersurfaces are then birational to quotients of $C_{(w_0,w_1,w_2)}\times
K_i$ by ${\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}/\ell{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}$. We list some examples found after a brief manual
search in Table 3. Of course most of these also occur in Table 1.
\begin{table}
\[ \begin{array}{|l|l| c | l | c|c|} \hline (w_0,w_1,w_2)
& (v_0,v_1,v_2,v_2) & \ell & (k_1,k_2,k_3,k_4,k_5) & d & \chi
\\ \hline \hline
(2,1,1) & (1,1,2,2) & 6 & (1,1,2,4,4) & 12 & -192 \\
& (1,2,3,6) & 12 & (1,1,4,6,12) & 24 & -312 \\
& (1,6,14,21) & 42 & (1,1,12,28,42) & 84 & -960 \\ \hline
(3,1,2) & (1,1,2,2) & 6 & (1,2,3,6,6) & 18 & -144 \\
& (1,2,3,6) & 12 & (1,2,6,9,18) & 36 & -228 \\
& (1,6,14,21) & 42 & (1,2,18,42,63) & 126 & -720 \\ \hline
(4,1,3) & (1,1,2,2) & 6 & (1,3,4,8,8) & 24 & -120 \\
& (1,2,3,6) & 12 & (1,3,8,12,24) & 48 & -192 \\
& (1,6,14,21) & 42 & (1,3,24,56,84) & 168 & -624 \\ \hline
(5,1,4) & (1,2,3,6) & 12 & (1,4,10,15,30) & 60 & -168 \\ \hline
(7,1,6) & (1,2,3,6) & 12 & (1,6,14,21,42) & 84 & -132 \\
& (1,6,14,21) & 42 & (1,6,42,98,147) & 294 & -480 \\ \hline
(5,2,3) & (1,1,2,2) & 6 & (2,3,5,10,10) & 30 & -72 \\
& (1,2,3,6) & 12 & (2,3,10,15,30) & 60 & -108 \\
& (1,6,14,21) & 42 & (2,3,30,70,105) & 210 & -384 \\ \hline
\end{array}
\]\caption{K3-fibered Calabi-Yau weighted hypersurfaces which are
also elliptic fibered, have constant modulus and are of Fermat type}
\end{table}
We now show by example that {\it a Calabi-Yau threefold can have two
different K3-fibrations with constant modulus}.
\begin{lemma}\label{theorem2} An appropriately choosen
weighted hypersurface of degree 12 in the weighted
projective space $\fP_{(2,2,1,1,6)}$
has two K3-fibrations with constant modulus; the two
different fibers are $\fP_{(1,1,4,6)}[12]$ and $\fP_{(1,1,1,3)}[6]$.
\end{lemma}
{\bf Proof:} Both of these fibrations can be constructed with the twist
map. For the first, we let ${\bgm}_3$ act on the product
$\fP_{(2,1,1)}[6]\times \fP_{(4,1,1,6)}[12]$; the twist map is onto a
hypersurface of degree 24 in $\fP_{(4,4,2,2,12)}$, which is the same thing
(as all weights are divisible by 2) as a hypersurface of degree 12 in
$\fP_{(2,2,1,1,6)}$. By Corollary \ref{cfiber}, we get a constant modulus
fibration with fiber $\fP_{(4,1,1,6)}[12]$. For the second, we let
${\bgm}_6$ act on the product $\fP_{(2,1,1)}[12]\times
\fP_{(1,1,1,3)}[6]$. The image under the twist map is a hypersurface of
degree 12 in $\fP_{(2,2,1,1,6)}$, and by choosing equations of the K3
surface and auxiliary curves appropriately, these two hypersurfaces
coincide. In particular it is true of the Fermat hypersurface. \hfill $\Box$ \vskip0.25cm } \def\cD{\ifmmode {\cal D
\subsubsection{A curious example}
Consider the ten K3 surfaces of Fermat type in the table in section
\ref{sK3}. For each, of weights $(k_1,k_2,k_3,k_4)$ and degree $d=\sum
k_i$, we can form the image under the twist map with $\fP_{(2,1,1)}$,
\[ \fP_{(2,1,1)}[2d] \times \fP_{(k_1,k_2,k_3,k_4)}[d] \longrightarrow} \def\sura{\twoheadrightarrow
\fP_{(1,1,2k_2,2k_3,2k_4)}[2d].\] For the cases 1, 4, and 9 in that list, the
resulting Calabi-Yau threefolds are the first three entries in Table
3. The analysis described there applies also to these cases, and the K3
fibration over $\fP^1$ has $2d$ singular fibers, which are the affine K3
singularities listed in Table 2 of part II of this paper.
Let {\bf X} be the type of bad fiber,
$e({\bf X})$ its Euler number. Then we have the formula for the Euler
number of the Calabi-Yau
\[ e(X) = (2-2d) \cdot 24 + 2d \cdot e({\bf X}) = 48 + 2d\cdot (e({\bf X})
- 24).\]
Now since $0<e({\bf X})<24$, we have $-24 < e({\bf X}) -24 <0$ which yields
the inequality for the Euler number of $X$:
\[ -48 d < e(X) - 48 = 2d(e({\bf X})-24) < 0.\]
Observe that ${\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}/d{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}$ is an automorphism group of the K3 $V_2$,
so by the results described above we have $d\leq 66$. Realizing that the case
with minimal known Euler number $-960$ is realized in this manner by taking
$V_2$ to be the K3 with $d=42$ (this is the third example in Table 3),
while there are cases $d=44, 66$, one could
imagine constructing a similar example with the $d=66$ case.
Next observe that if $g$ denotes the generator of $\hbox{Aut}} \def\Im{\hbox{Im}} \def\mod{\hbox{mod}(V_2)$, fixing $x\in
\Delta\subset} \def\nni{\supset} \def\und{\underline \fP^1$ in the discriminant of the fibration, formula (\ref{8})
shows the contribution of the singular fiber $X_x$ at $x\in \Delta$ is
determined by the action of $g$ on the holomorphic two-form. In each case
of the fibrations described above, for each $x_i\in \Delta,\ a_i={1\over d}$,
and we have the equality
\[ \sum_{i=1}^{2d} a_i = 2d {1\over d} = 2 = c_1(\fP^1),\]
which is the necessary condition described at the beginning of this
paper. For the case $d=66$ this would require 132 singular fibers, each
giving a contribution of ${1\over 66}$. If so, we can calculate what its
Euler number would be. Since it
has 132 singular fibers which are the affine surface $t_1^{11} +t_2^3
+t_4^2 = 0 $ which has Milnor number $\mu = (11-1)(3-1)(2-1)=20$, each
singular fiber has Euler number 4, and our formula for the Euler number
yields:
\[ e(X) = (2-132) \cdot 24 + 132\cdot 4 = -2592.\]
Is it possible to contruct such an
example?
Consider the image of the twist map
\[ \fP_{(2,1,1)}[132] \times \fP_{(1,6,22,33)}[66] \longrightarrow} \def\sura{\twoheadrightarrow
\fP_{(1,1,12,44,66)}[132].\]
If we take $V_1,\ V_2$ as Fermat hypersurfaces, then the image is
the weighted threefold given by the equation
\[ X = \{ z_1^{132} + z_2^{132} + t_1^{11} +t_2^3 +t_4^2 = 0 \} \subset} \def\nni{\supset} \def\und{\underline
\fP_{(1,1,12,44,66)}. \]
Suppose this threefold did allow a fibration with fibers the resolved K3
surfaces discussed in section \ref{exoticsurface}.
It is easy to see that we have
singular fibers at the 132 points which are the zeroes of $z_1^{132} +
z_2^{132}$, and that the fibers $\{z_1=0\}\cap X$ and $\{z_2=0\}\cap X$ are
both smooth, so that it looks as if we could in fact construct such a
fibration. However, as $132 > 124=\sum k_i$, the threefold $X$ does not
satisfy the sufficient condition to be Calabi-Yau, and our arguments above
(in particular Lemma \ref{l3.4})
do not apply directly. In fact, assuming the necessary condition (\ref{4})
is also sufficient, such a fibration does {\it not} exist,
which can be seen as follows. There is no birational model of $X$ which is
Calabi-Yau: the geometric genus
$h^{3,0}$ is 9, not 1 as would be the case if $X$ were Calabi-Yau. Indeed,
since the geometric genus is a birational invariant, this is
just the number of sections of $\cO(132-124)=\cO(8)$, and this is the
number of monomials in the first two variables of weight 8, of which there
are 9. However, using the method of the proof of Lemma \ref{l3.4}, the
reader may verify that in fact such a fibration does exist (first the fiber
is the non-Calabi-Yau surface $\{x^2+y^3+z^{11}+w^{66}=0\}\subset} \def\nni{\supset} \def\und{\underline
\fP_{(1,6,22,33)}$ of section 3.3.3, which is birationally modified as in
that section over the base of the fibration). Consequntly, condition
(\ref{4}) is not sufficient.
We note that we can also display
$X$ as an elliptic fibration, and for these we have the necessary and
sufficient condition (\ref{6}), which utilizing the Weierstra\ss\ form says
that
\[ -12 K_Y = \Delta,\]
where $\Delta$ is the discriminant locus (counted with appropriate
multiplicities). The variety $X$ is the image of the twist map
\[ \Phi:\fP_{(11,1,1,12)}[132] \times \fP_{(1,2,3)}[6] \longrightarrow} \def\sura{\twoheadrightarrow
\fP_{(1,1,12,44,66)}[132].\]
The discriminant is the total transform of
\[ \Sigma = \{x_1^{132}+x_2^{132}+x_3^{11} =0 \} \subset} \def\nni{\supset} \def\und{\underline \fP_{(1,1,12)};\]
projecting onto $\fP_{(1,1)}$ displays the {\it proper transform} of this,
which we denote also by $\Sigma$, as an 11 to 1 cover, totally
branched at the 132rd roots of $-1$, so it has Euler number
\[ e(\Sigma) = 11(2-132) + 132 = -1298.\]
We can determine its class in ${\rm H}^2(\fF_{12},{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P})$ as follows: from the fact
that it is an 11 to 1 cover, it intersects a fiber in 11 points, so
$\Sigma \sim 11 C_0 + b F$, where $C_0$ denotes the (class of) a section of
positive self-intersection and $F$ denotes the class of a fiber. On the
other hand, the first Chern class of the resolution $\fF_{12}$ of
$\fP_{(1,1,12)}$ is
\[ c_1(\fF_{12}) = 2 C_0 - 10 F,\]
and applying adjuction to $\Sigma$ on $\fF_{12}$,
which is a smooth curve with Euler number we just calculated, one can
determine its class in ${\rm H}^2(\fF_{12},{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P})$. The result is
\[ \Sigma \sim 11 C_0 \sim 11( C_{\infty} +12F),\]
while for the exceptional curve $C_{\infty}$, we have fiber types II again,
so
\[ \Delta = 2\Sigma + 2 C_{\infty} = 24 C_{\infty} + 264 F,\]
and one sees that the sufficient condition (\ref{4}) is not
satisfied, as $-12K_{\fF_{12}} = 24 C_{\infty} + 168 F$.
It remains an open problem whether it is possible to construct such a K3
fibration; if possible, it would enlarge the range of possible Euler
numbers of Calabi-Yau threefolds (and is therefore quite unlikely).
\subsubsection{Birational fibrations}
It is, however, possible to construct K3 fibrations which have the K3
surface of section \ref{exoticsurface} as fiber, which we now explain.
There are examples of weighted hypersurfaces which, after resolution of
singularities, are not fibrations, but still are, in the class of
Calabi-Yau threefolds, birational to such a Calabi-Yau. These examples come
from the exotic surface example of section \ref{exoticsurface}. Indeed,
suppose we want a twist map
\[ \Phi : \fP_{(w_0,w_1,w_2)}[66\cdot w_0]\times \fP_{(1,6,22,33)}[66] \longrightarrow} \def\sura{\twoheadrightarrow
\fP_{(w_1,w_2,w_0\cdot 6,w_0\cdot 22,w_0\cdot 33)}[66\cdot w_0]\]
to have an image which satisfies the sufficient condition $d = \sum k_i$
for it to be a Calabi-Yau threefold.
Then writing down what the weights are,
we get an equation $w_1+w_2 +w_0(6+22+33) \stackrel{!}{=}
66w_0$, which means that we
require $w_1+w_2 =5w_0$. The easiest solution to this is:
$w_0=1, (w_1,w_2) = (2,3)$ or
$(1,4)$. The image projective spaces is then $\fP_{(2,3,6,22,33)}$, and
there are indeed Calabi-Yaus (of degree 66) in this space, as well as in
$\fP_{(1,4,6,22,33)}$. They have Euler numbers $-240$ and $-300$,
respectively.
They do not posess a fibration {\it a priori}, because the base locus
of the projection (which one gets by setting the
first two coordinates =0) is not part of the singular
locus (recall that in Lemma \ref{l3.4} we required $w_0>1$ to get a
fibration). However, one can blow it up (upon which the surface is temporarily
no longer C-Y) and then in each fiber do the blowing down process we have
described in section \ref{exoticsurface}.
After this is done, we do get a C-Y with fibration by those exotic surfaces.
\section{Applications in Physics}
In this section we breifly describe some of the applications of the twist
map to physical dualities, which was the original source of motivation for
the present investigation. It is intended for the non-expert, and we just
try to explain the physical interpretation of the geometry, without going
into any details.
\subsection{The physical theories and their moduli spaces}
We are interested in {\it superstring theories}. These are theories about
how a string (a smooth image of the circle) moves in Minkowski space; its
vibrations give rise to all elementary particles. More precisely, consider
a string moving in some Minkowski space $M^{1,d}$; it traces out with time
a {\it world sheet}, and this is then an embedded Riemann surface in
$M^{1,d}$. The physical theory one is interested in is a {\it
superconformal field theory} on that Riemann surface; the adjective
superconformal refers to the symmetry group of the equations of motion. It
is well-known that for this symmetry to hold, the dimension is restricted
to $d=9$ (ten-dimensional Minkowski space), and it is also known that there
are five consistent theories of this type: Types I, IIA, IIB,
Het$_{E_8\times E_8}$, Het$_{SO(32)}$. The first is the theory which
contains open strings and has only $N=1$ supersymmetry, while the others
are theories of closed strings and have $N=2$ supersymmetry (on the world
sheet). That is, one has a consistent supersymmetric
quantum field theory on the world
sheet of the string. (There is also an issue of space-time supersymmetry,
but we neglect that here). The type II theories have at most abelian gauge
groups (at least in ten dimensions, see \cite{SS2}, Chapter 14 for a
discussion of this issue upon compactification),
which is why they seemed completely uninteresting for many years,
while the heterotic strings, which have only left-moving supersymmetries,
have the two gauge groups of rank 16 (which arises from the fact that the
consistent dimension for the {\it bosonic}, or non-supersymmetric string,
is 26, and the $16=26-10$ remaining dimensions are compactified to the
maximal torus of a Lie group, the gauge group) listed above.
The superstring theories are all {\it perturbative} in essence, which means
(contrary to general relativity) they are small perturbations of certain
vacuum solutions.
These theories are consistent, but phenomenologically uninteresting, as
in reality one does not observe ten flat dimensions.
To aleviate this, one can compactify
six of the dimensions, and making the size of the compact factor small
enough, it will be invisible to us yielding a phenomenologically more
acceptable theory. The space time in which the world sheet
resides is then $M^{1,3} \times X$ for some compact manifold $X$. In
order to assure superconformal symmetry here also, it is required that $X$
be Calabi-Yau\footnote{the super- invariance implies K\"ahler, the
conformal invariance implies Ricci flatness}.
Such a manifold has two types of moduli: complex structures
and K\"ahler forms. These moduli turn out to be moduli of the physical
theory also, that is, any $X$ (whatever its moduli) gives a consistent
compactification, and in this way, moduli spaces enter also in the
physical theory. These correspond in a sense to ``flat directions of the
potential'', in other words change only the explicit form of the
Lagrangian, not the equations of motion. The ``physical moduli space'' will
in general consist of these geometric moduli in addition to others.
In order to better understand
these compactifcations, one often compactifies fewer than six dimensions,
an example of which we discuss next.
\subsection{Toy model: IIA(K3) $\longleftrightarrow$ Het$_{E_8\times
E_8}$(T$^4$)}
The notation of the title is meant to indicate that one starts with one of
the five theories above and compactifies on the Calabi-Yau manifold in
parenthesis, in this case of dimension four, leaving a six-dimensional
Minkowski space left as ``space-time''. The arrow indicates the so-called
{\it duality}; this means that the {\it non-perturbative} theory, of which
the supersting theory is an approximation (perhaps it would be better to
say the superstring theory is a certain {\it limit}), is {\it the same} on both
sides. Think of this as meaning there is an underlying theory, with certain
moduli, and at certain moduli points one can approximate the theory by the
two different superstring theories. It is quite involved to list what this
implies physically, but one thing is certain: if this is the case, then
both sides must necessarily have the same moduli space, and this can be
verified and even understood.
The moduli
space in question is given by
\begin{equation}\label{modsp}
{\cal M} = {\rm SO}(4,20;{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}) {\large \backslash} {\rm SO}(4,20) {\large /}
{\rm SO(4)}\times {\rm SO(20)}
\end{equation}
and the heterotic/type II duality reduces in this context to
giving two different interpretations to this moduli space. The group
${\rm SO}(4,20;{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P})$ is the discrete group preserving a particular lattice
in $\fR^{24}$, which is described below.
The symmetric space on the right-hand side of (\ref{modsp})
can be described as the space
of 4-dimensional subspaces $V$ of the
space $\fR^{(4,20)}$, by which we mean $\fR^{24}$
endowed with a metric of signature (4,20),
on which the metric is positiv definite.
The space $\fR^{(4,20)}$ contains a (unique) selfdual even integral lattice,
which we denote by ${\Gamma}^{(4,20)} \subset \fR^{(4,20)}$, and which
has the same signature (in the above notations, ${\Gamma}^{(4,20)} \cong
U^{\oplus 4}\oplus E_8^{\oplus 2}$).
The moduli space ${\cal M}$ then can be described
as the space of all $V$ up to automorphisms of the lattice
${\Gamma}^{(4,20)}$.
Now, in the heterotic string interpretation of this space the subspace
$V$ and its orthogonal complement $V^{\perp}$ are associated to
gauge fields
describing the Yang-Mills structure of the string. Generically, the
non-abelian
gauge group $E_8\times E_8$ of the heterotic string is broken to the
abelian group $U(1)^{16}$ (see for example \cite{Aspin}, section 3.5 and
page 468),
i.e., none of the non-abelian gauge fields persist upon compactification. Of
special significance are those points in the moduli space
for which $V^{\perp}$ contains nonzero points of the lattice
${\Gamma}^{(4,20)}$. If $V^{\perp}$ contains such nonzero points
for which the normalized length $d_P = -\frac{1}{2}<P,P>$
equals unity then physical considerations lead to an enhancement
of the rank of the Yang-Mills group beyond the rank encountered
at a generic point of the moduli space. So what is varying here in the
moduli space is the surviving non-abelian gauge group.
The second, type II string, interpretation of the moduli space
(\ref{modsp}) is obtained by viewing $\fR^{(4,20)}$ as the
real cohomology ${\rm H}^*$ of a K3 surface $S$. It is known that type IIA
compactifed on a K3 surface is equivalent to {\it a superconformal
non-linear sigma models on that K3 surface}. Its moduli space can be
described as follows, see \cite{Aspin}
\S 3 for details on this aspect. From the decomposition
\[ { SO(4,20) \over S(O(4) \times O(20))} \cong {SO(3,19) \over
S(O(3)\times O(19))} \times \fR^{22} \times \fR_+,\]
this has the following interpretation: the first factor is the space of
{\it Einstein} metrics on a K3 surface, the second factor is the moduli
space of the so-called $B$-field and the final factor is the size.
It remains to determine the meaning of the classes of norm $-2$ occuring
above. But it is well-known how this occurs. Referring back to section
3.4.2 (and letting our K3 surface be denoted by $X$ as there), we have the
Picard lattice $S_X$ and the transcendental lattice $T_X$. The Picard
lattice is spanned by the K\"ahler class and by rational $-2$ curves,
which correspond to a generic K3 surface aquiring ordinary double points
and then resolving these to yield a smooth K3 surface. In this respect it
is important to recall that the number of moduli preserving such a Picard
lattice is $20-\rho$, see for example \cite{book}, Proposition 2.3.2. For a
generic {\it algebraic} surface, there are 19 moduli and the Picard number
$\rho$ (the rank of the Picard group) is 1, while for a K3 surface with
$\rho >1$, there are only $20-\rho$ moduli of K3 surfaces which are
deformations of the given one and which have isomorphic Picard lattice.
The relation in this case to the heterotic theory was observed by Witten in
\cite{ew95}. Namely, if $S$ aquires a
singularity of any of the A,D, or E types, a configuration of
rational curves whose dual graph is the Dynkin diagram of one of the groups
of type A, D or E, collapses
to a point (these are the vanishing cycles).
Such rational curves have selfintersection $-2$ and thus
correspond precisely to the vectors encountered above in the
heterotic description.
\subsection{String dualities in $D=4$}
Next we consider the compactification of Type IIA on a Calabi-Yau threefold
(down to $D=4$) and the compactification of the heterotic string on a
product $S\times T^2$, where $S$ is a K3 surface. Let us think of this in
the complex category and make the simplifying assumption that $S$ is
elliptically fibered (this is an 18-dimensional family, as compared with the
19-dimensional family of K3 surfaces with some fixed polarization) and that
the Calabi-Yau threefold $X$ is fibered by K3 surfaces. Then, we have two
fibrations:
\[ S \longrightarrow} \def\sura{\twoheadrightarrow \fP^1, \quad \quad X\longrightarrow} \def\sura{\twoheadrightarrow \fP^1,\]
and combining the first with $T^2$, we have two fibrations, one of $X$ and
one of $S\times T^2$, both onto $\fP^1$. It is natural to think of the
above duality for a fixed $t\in \fP^1$, and applying the duality {\it
fiberwise}. There is a physical argument for this, the so-called {\it
adiabatic limit}. Note however, that if $S$ is one of our K3 surfaces and
$X$ one of our Calabi-Yau threefolds, for which the modulus of the fiber is
{\it constant}, then there is no argument at all necessary: applying the
duality above fiberwise implies that the gauge group of the heterotic
theory is just the group whose weight lattice is the Picard group of the K3
fiber on $X$.
For these $D=4$ theories, the heterotic side is quite complicated; there
are vector bundles (background fields) on the K3 surface involved, and a
Higgsing process. Nonetheless, our constructions gives us a good idea of
what the gauge group of a heterotic dual theory would be. In \cite{Alz},
there are chains of such dual theories. We give four examples, three of
which are in \cite{Alz}, the other of which is
considered in \cite{Alz2}. We start
with the K3 surfaces listed as \# 5,7,8 and 9 in section 3.3.2. We then form
the product with the curve $C$ with weights $(2,1,1)$ and $\ell = 18, 24,
20$ and 42 in the respective cases. The image of the twist map is in these
four cases a weighted hypersurface with weights $(1,1,4,12,18),\
(1,1,6,16,24),\ (1,1,8,10,20),\ (1,1,12,28,42)$ of degrees $36,\ 48,\ 40,\
84$. Looking at the table in section 3.3.2, we see that the singular fibers
I$_0^*$,\ IV$^*$,\ III$^*$ and II$^*$
of the elliptic fibrations of the K3 surfaces (the fibers of the
fibration) have dual graphs which are the extended Dynkin diagrams of the
types $D_4,\ E_6,\ E_7$ and $E_8$, respectively, yielding as gauge
groups of the heterotic duals $SO(8),\ E_6,\ E_7$ and $E_8$, respectively.
For the three cases which occur in \cite{Alz}, these are (up to abelian
factors) indeed the gauge groups of the heterotic dual theories, and for
the remaining case this is verified in \cite{Alz2}, p.\ 133.
As explained in \cite{Asp2}, the moduli of the two physical theories are
described in more detail as follows. The type IIA string compactified on a
Calabi-Yau threefold $X$ has the following moduli:
\begin{enumerate} \item The dilaton (which governs the string coupling
constant) and the axion which together form a complex scalar $\Phi_{\rm
IIA}$;
\item A metric which is determined by the complex structure of $X$ and
the cohomology class of a K\"ahler form;
\item A skew field $B \in {\rm H}^2(X,\fR)/{\rm H}^2(X,{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P})$;
\item So-called Ramond-Ramond fields $R\in {\rm H}^{odd}(X,\fR)/{\rm H}^{odd}(X,{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P})$.
\end{enumerate}
These moduli split into two types, the so-called {\it vector} moduli and
the {\it hypermultiplet} moduli, as follows:
\begin{itemize}\item[V] The K\"ahler form and the $B$-field together form
the well-known ``complexified K\"ahler form'' used in mirror symmetry;
these moduli together form a moduli space $\cM_V$ which is a special
K\"ahler manifold.
\item[H] The complex scalar $\Phi_{\rm IIA}$, the Ramond-Ramond fields
$R$ and the complex structure of $X$ form the moduli space $\cM_H$ of
hypermultiplet moduli; this is a quaternionic K\"ahler manifold.
\end{itemize}
The above descriptions are not valid in the complete quantum theory, but
rather only for certain approximations: $\cM_V$ is valid only in the
``large radius limit'' of $X$, $\cM_H$ is valid only near the
weakly-coupled limit $\Phi_{\rm IIA}\longrightarrow} \def\sura{\twoheadrightarrow -\infty$.
The heterotic string compactified on a product of a K3 surface $S_H$ and an
elliptic curve $E_H$, $S_H\times E_H$ (which is K\"ahler Ricci-flat) has
the following moduli:
\begin{enumerate} \item The dilaton (which governs the string coupling) and
the axion, which together form a complex scalar $\Phi_{\rm Het}$.
\item A Ricci-flat metric on the product $S_H\times E_H$.
\item A skew-field $B\in {\rm H}^2(S_H\times E_H,\fR)/{\rm H}^2(S_H\times
E_H,{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P})$.
\item A $G$-bundle on $S_H\times E_H$ with a connection satisfying the
Yang-Mills equations, where $G$ is the (unbroken) gauge group of the
heterotic string, either ${\rm Spin}(32)/{\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}_2$ or $E_8\times E_8$.
\end{enumerate}
Once again, these moduli split into two types, vector and
hypermultiplet. Assume that the $G$ bundle is the product of a
$G_S$-bundle over $S_H$ and a $G_E$-bundle over $E_H$, where $G_S\times
G_E\subset G$ is a subgroup. Then these types can be described as follows:
\begin{itemize}\item[V] The scalar $\Phi_{\rm Het}$, the moduli of the
$G_E$-bundle over $E_H$, the metric on $E_H$ and the $B$-field on $E_H$
form the vector muliplet space $\hbox{\script M}_V$.
\item[H] The moduli of the $G_S$-bundle on $S_H$, the metric on $S_H$ and
the $B$-field on $S_H$ form the hypermultiplet moduli space $\hbox{\script M}_H$.
\end{itemize}
Again, these descriptions are only valid in certain limits: $\hbox{\script M}_V$ when
$\Phi_{\rm Het} \longrightarrow} \def\sura{\twoheadrightarrow -\infty$ and the area of $E_H$ is large, and $\hbox{\script M}_H$
when the volume of $S_H$ is large. Duality here means essentially matching
these moduli spaces in the two cases. The match of vector moduli is
described above, and the match of hypermultiplet moduli leads to quite
interesting mathematical constructs, for example intermediate Jacobians,
Prym varieties and Deligne cohomology, see \cite{CD}.
Perhaps the most fascinating aspect of this is determining the K3 surface
explicitly in terms of the Calabi-Yau threefold $X$; for this one considers
a stable degeneration of $X$ into the union of two generalized
Fano-threefolds $X_1 \cup X_2$, and the K3 surface for the heterotic
compactification is the intersection $X_1\cap X_2$.
\subsection{Conifold transitions}
A second problem which is illuminated by our construction is
the heterotic structure of the so-called conifold transition
between Calabi-Yau manifolds. Such transitions are given by the following
construction. Allow a smooth Calabi-Yau threefold $X_t$ depending on a
parameter $t$ to aquire a certain
number of ordinary double points at $t=0$;
let $X^*(=X_0)$ denote the singular space. Each
of the ordinary double points can be resolved by a small resolution; we
assume that there is at least one of these which is projective (i.e.,
K\"ahler), and let $X^s$ denote such a resolution.
Schematically we have the following situation:
\[ X_t \longleftrightarrow} \def\ra{\rightarrow} \def\Ra{\Rightarrow X^* \longleftrightarrow} \def\ra{\rightarrow} \def\Ra{\Rightarrow X^s.\]
The Hodge numbers of $X_t \ (t\neq 0)$
and $X^s$ are related as follows:
\[h^{2,1}(X^s) = h^{2,1}(X_t)-(P-R),\quad \quad h^{1,1}(X^s) = h^{1,1}(X_t)
+R,\]
where $P$ denotes the number of nodes and $R$ denotes the number of
relations between the corresponding vanishing cycles.
Although this transition passes through a singular space $X^*$, Strominger
has shown \cite{strom} that the physics remains smooth. What happens is
something quite similar to what occurs in the work of Seiberg and Witten: a
massive particle (in this case a black hole) gets massless at the moduli
point of $X^*$ (in the physical theory one always makes a low-energy
approximation, and all massive particles are so heavy that they do not
influence the physics; accordingly they are integrated out of the
Lagrangian), and passing to $X^s$ amounts to the new theory with an
additional massless particle.
\subsubsection{Splittings}
One way of describing such transitions is by means of {\it splittings},
which are described in terms of complete interesections in products of
weighted projective spaces. As an example of this, consider first a
transversal weighted hypersurface $\fP_{(k_1,k_1,k_2,k_3,k_4)}[d]$, where
$d= 2k_1+k_2+k_3+k_4$, and consider the following threefold in the product
$\fP_{(1,1)}\times \fP_{(k_1,k_1,k_2,k_3,k_4)}$:
\[ X_0:= \left\{ \begin{array}{c} p_1(u,y)=u_0Q(y)+u_1R(y)=0 \\ p_2(u,y) =
u_0S(y)+u_1T(y)=0
\end{array} \right\} \subset} \def\nni{\supset} \def\und{\underline \fP_{(1,1)}\times
\fP_{(k_1,k_1,k_2,k_3,k_4)}.\]
Schematically this is abbreviated with the following notation: \[X_0 \in
\begin{array}{l} \fP_{(1,1)} \\ \fP_{(k_1,k_1,k_2,k_3,k_4)}
\end{array} \left[ \begin{array}{cc} 1 & 1 \\ a\cdot k_1 & d-a\cdot k_1
\end{array} \right],\]
where $a\cdot k_1=deg(Q)=deg(R)$ and $d-a\cdot k_1 = deg(S)=deg(T)$. On the
other hand, consider the determinental variety
\[ X^* := \{ Q(y)T(y) - R(y)S(y) =0 \} \subset} \def\nni{\supset} \def\und{\underline \fP_{(k_1,k_1,k_2,k_3,k_4)}.\]
Clearly $X^*$ is singular for generic choices of $Q,\ R,\ S$ and $T$ when
$Q=R=S=T=0$, which means that $X^*$ generically has isolated singularities,
which one can check are ordinary double points. Furthermore, mapping
$\fP_{(k_1,k_1,k_2,k_3,k_4)}$ into the product $\fP_{(1,1)}\times
\fP_{(k_1,k_1,k_2,k_3,k_4)}$ in the obvious way, it is clear that $X^*$
maps to $X_0$: write the equation defining $X_0$ as \[P(u,y) =
(u_0,u_1)\left( \begin{array}{cc} Q(y) & R(y) \\ S(y) & T(y)
\end{array}\right) = (u_0,u_1)\Pi(y) =0,\]
with a $2\times 2$ matrix $\Pi$. Then $y\in X^* \iff \det(\Pi(y))=0 \iff
\exists_{(u_0,u_1)}$ with $(u_0,u_1)$ is in the kernel of $\Pi(y)$ $\iff$
$P(u,y)=0 \iff (u,y)\in X_0$. The singular $X^*$ can be deformed by adding
some multiple of a transversal polynomial, i.e., by setting
\[ X_t := \{ t_0( Q(y)T(y) - R(y)S(y) ) + t_1 p_{trans}(y) =0 \} \subset} \def\nni{\supset} \def\und{\underline
\fP_{(k_1,k_1,k_2,k_3,k_4)}\quad (t=(t_0,t_1)\in \fP^1) ,\]
and we are in the situation mentioned above: the smooth $X_t$ aquires
singularities of the desired type (ordinary double points) at $t_1=0$, and
this singular $X_0$ can be given a small resolution $X^s$. Once again, we
{\it schematically} describe this process by the symbols
\[ \fP_{(k_1,k_1,k_2,k_3,k_4)}[d] \longleftrightarrow} \def\ra{\rightarrow} \def\Ra{\Rightarrow \begin{array}{l} \fP_{(1,1)} \\
\fP_{(k_1,k_1,k_2,k_3,k_4)}
\end{array} \left[ \begin{array}{cc} 1 & 1 \\ a\cdot k_1 & d-a\cdot k_1
\end{array} \right].\]
\subsubsection{K3 fibrations}
It was shown in \cite{ls95} that such transitions can be
constructed between K3-fibered Calabi-Yau manifolds, an example
being provided by the transition
\begin{equation}\label{hetsplit}
\fP_{(1,1,2,4,4)}[12]^{(5,101)} ~\longleftrightarrow ~
\begin{array}{l}\fP_{(1,1)}\\ \fP_{(4,4,1,1,2)}
\end{array}
\left[\begin{array}{cc} 1 & 1 \\ 4 & 8 \\
\end{array} \right]^{(6,70)},
\end{equation}
where the notation on the right denotes a complete intersection manifold
of codimension two defined by two polynomials of bi-degree
(1,4) and (1,8) respectively and the superscripts indicate the
Hodge numbers $(h^{(1,1)},h^{(2,1)})$.
Note that the smooth hypersurface on the left hand side is the first
example in Table 3, so that the Fermat hypersurface is in fact the image of
an appropriate twist map. In this case we have $\hbox{deg}} \def\Pic{\hbox{Pic}} \def\Jac{\hbox{Jac}(Q)=\hbox{deg}} \def\Pic{\hbox{Pic}} \def\Jac{\hbox{Jac}(R)=4$ and
$\hbox{deg}} \def\Pic{\hbox{Pic}} \def\Jac{\hbox{Jac}(S)= \hbox{deg}} \def\Pic{\hbox{Pic}} \def\Jac{\hbox{Jac}(T)=8$, hence $Q=R=S=T$ has (after rescalings) 32
solutions. Hence the singular $X^*$ above has $P=32$ and $R=1$, which
explains the change in Hodge numbers explicitly from this point of view.
To see the K3-fibration on the right hand side, one considers the sections
$\lambda_0 y_0-\lambda_1y_1=0$; an easy calculation shows that these are K3 surfaces
in the family of complete intersections in $\fP_{(1,1)}\times
\fP_{(1,1,2,2)}$ of degrees $(1,2)$ and $(1,4)$.
\subsubsection{A generalized twist map}
To see that the right hand side above can also be realized as a {\it
constant modulus} fibration, we can generalize the twist map to this
situation. Define the following rational map:
\begin{eqnarray}\label{gentwist} \Phi: \fP_{(w_0,\ldots,w_n)} \times
\fP_{(1,1)} \times \fP_{(v_0,\ldots,v_m)} & \longrightarrow} \def\sura{\twoheadrightarrow & \fP_{(1,1)} \times
\fP_{(w_1v_0,\ldots, w_nv_0,w_0v_1,\ldots, w_0v_m)} \\
((x_0,\ldots, x_n),(u_0,u_1),(y_0,\ldots, y_m)) & \mapsto &
((u_0,u_1),(y_0^{w_1/w_0}x_1,\ldots,
y_0^{w_n/w_0}x_n,x_0^{v_1/v_0}y_1,\ldots, x_0^{v_m/v_0}y_m)).\nonumber
\end{eqnarray}
Let the subvarieties $V_1,\ V_2$ be defined as follows:
$V_1=\{x_0^{\mu}+p(x_1,\ldots, x_n) =0 \} \subset} \def\nni{\supset} \def\und{\underline
\fP_{(w_0,\ldots, w_n)};$
\begin{equation}\label{vees} \quad V_2 = \left\{ \begin{array}{l}
p_1({\bf u},{\bf y})= u_0y_1 + u_1\cdot p_{11}(y_1,\ldots, y_m) =0 \\
p_2({\bf u},{\bf y})= u_0(y_0^{\mu}+p_{20}(y_1,\ldots,y_m))+
u_1y_1^{\nu-1} =0
\end{array} \right. \subset} \def\nni{\supset} \def\und{\underline \fP_{(1,1)}\times \fP_{(v_0,\ldots, v_m)}.
\end{equation}
This complete intersection has bidegrees $\left[\begin{array}{cc}1 & 1 \\
v_1 & d-v_1
\end{array} \right]$,
where $d=\sum_0^m v_i$; $p_{11}$ has degree $v_1$, and
the degree of $p_{20}$ is $\hbox{deg}} \def\Pic{\hbox{Pic}} \def\Jac{\hbox{Jac}(p_{20}) = (\hbox{deg}} \def\Pic{\hbox{Pic}} \def\Jac{\hbox{Jac}( y_1))(\nu-1) = v_1(\nu-1) =
\hbox{deg}} \def\Pic{\hbox{Pic}} \def\Jac{\hbox{Jac} (y_0) \cdot \mu = v_0\cdot \mu = d-v_1$, hence we have the relation
among the various weights:
\[ \mu = {v_1(\nu-1)\over v_0} = {d-v_1 \over v_0}.\]
Clearly
$V_1$ is invariant under the obvious action of ${\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}_{\mu}$. We claim that
$V_2$ is invariant under the following action of ${\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}_{\nu}$:
\[ ({\bf u,y}) \mapsto ((\eta^{\nu-1}u_0,u_1),(\eta y_0,y_1, \ldots,
y_m))\]
for a generator $\eta$ of ${\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}_{\nu}$. Indeed, the first equation defining
$V_2$ is invariant, while the second gets multiplied by a factor of
$\eta^{\nu-1}$, hence the zero locus is invariant. Set $\ell := {\rm
gcd}(\mu,\nu)$, then we get an action of ${\bf Z}} \def\fR{{\bf R}} \def\fC{{\bf C}} \def\fP{{\bf P}_{\ell}$ on the product
space as follows:
\[ (\zeta,{\bf x,u,y}) \mapsto ((\zeta^{\nu} x_0,x_1,\ldots,
x_n),(\zeta^{\mu\cdot (\nu-1)} u_0, u_1),(\zeta^{\mu} y_0, y_1, \ldots,
y_m)).\]
Finally, let $X$ be defined in the product \[ \fP_{(1,1)} \times
\fP_{(w_1v_0,\ldots, w_nv_0,w_0v_1,\ldots, w_0v_m)} \hbox{ with coordinates
} ((u_0,u_1),(z_1,\ldots, z_n, t_1,\ldots, t_m))\] by
\begin{equation}\label{eks} X = \left\{ \begin{array}{l} u_0\cdot t_1
+u_1p_{11}(t_1,\ldots,t_m) =0 \\
u_0(p_{20}(t_1,\ldots, t_m)-p(z_1,\ldots, z_n)) ) + u_1\cdot t_1^{\nu-1} =0
\end{array} \right. .
\end{equation}
Then a calculation shows that $\Phi(V_1\times V_2) \subset} \def\nni{\supset} \def\und{\underline X$,
and that this displays $X$ rationally as a quotient of $V_1\times V_2$ by a
${\bgm}_{\ell}$-operation. Indeed, the first equation defining $X$, in
terms of the coordinates $\bf x,y$, is
\[ u_0 \cdot x_0^{v_1/v_0}\cdot y_1 + u_1\cdot (x_0^{1/v_0})^{deg(p_{11})}
\cdot p_{11}(y_1,\ldots,y_m) = x_0^{v_1/v_0}(u_0 y_1 + u_1
p_{11}(y_1,\ldots, y_m)),\]
which clearly vanishes for ${\bf (u,y)} \in V_2$. A similar
calculation for the second equation is
\[ u_0\left( \left(x_0^{1/v_0}\right)^{v_0\cdot \mu} \cdot
p_{20}(y_1,\ldots, y_m) - \left( y_0^{1/w_0}\right)^{\mu \cdot w_0} \cdot
p(x_1,\ldots, x_n) \right) + u_1\left( x_0^{v_1/v_0}\right)^{\nu-1} \cdot
y_1^{\nu-1} ,\]
which again clearly vanishes for ${\bf (x,u,y)} \in V_1\times V_2$.
\subsubsection{Gauge groups}
Let us apply this to determine the gauge groups on both sides of
(\ref{hetsplit}). For convenience we stick to Fermat polynomials. On the left
hand side we have a constant modulus K3-fibration with fiber the K3 surface
occuring in the first line of the table in section 3.3.2, with six singular
fibers of type IV. Since each of these corresponds to a $A_2$, we get for
the Picard group of the K3 a lattice of type $A_2^6 \oplus H$, where $H$ is
a sum of hyperbolic and abelian factors. As we have already explained,
since the Calabi-Yau threefold is the quotient of the product by a group,
and the action of the group on the K3-surface is of non-Nikulin type, it
preserves the Picard lattice, hence this is, up to abelian factors, the
gauge group of the theory compactified on that Calabi-Yau.
To do the same for the right hand
side, we must first study the K3 fiber, which is the complete interesection
of type $(1,2)$, $(1,4)$ in $\fP_{(1,1)}\times \fP_{(1,1,2,2)}$. Recall
that we can determine the Picard group from the elliptic fibration; in this
case we first describe this K3 as the image of a twist map. This is given
by the above constuction, setting $(v_0,v_1,v_2)=(1,1,1),\ (w_0,w_1,w_2)
=(2,1,1)$, $p(x_1,x_2)=x_1^4+x_2^4$, $p_{11}(y_1,y_2)=y_2$ and
$p_{20}(y_1,y_2) = y_1^2$. In other words, let $V_1 = \{
x_0^2+x_1^4+x_2^4=0\}\subset} \def\nni{\supset} \def\und{\underline \fP_{(2,1,1)}$ (so $\mu =2$),
and let $V_2$ be the elliptic curve defined as follows:
\[ V_2 = \left\{ \begin{array}{l} u_0\cdot y_1+u_1\cdot y_2 =0 \\
u_0(y_0^2+y_1^2) + u_1\cdot y_1^2 =0
\end{array} \right. \subset} \def\nni{\supset} \def\und{\underline \fP_{(1,1)}\times \fP_{(1,1,1)}.\]
The image of $V_1\times V_2$ under the twist map is the complete
intersection $W$ defined as follows
\[ W = \left\{ \begin{array}{l} u_0t_1+u_1t_2 =0\\
u_0(t_1^2-(z_1^4+z_2^4))+u_1\cdot t_1^2 =0
\end{array} \right. \subset} \def\nni{\supset} \def\und{\underline \fP_{(1,1)}\times \fP_{(1,1,2,2)}.\]
In this case, $\nu=3$ and $\ell =6$, so we have a group of order six acting
on the product, with a subgroup of order 3 acting on the elliptic
curve. The singular fibers of the fibration are hence of types {\bf IV} or
{\bf IV$^*$}. It is easy to see that at the four zeros of $p(x_1,x_2)$, we
have singular fibers of type ${\bf IV}$. The reader may check that there
is, in addition, a singular fiber of type ${\bf IV^*}$. Hence, in this case
the Picard group is $A_2^4\oplus E_6$.
As we have described above, this is the same as a determinental
hypersurface in $\fP_{(1,1,2,2)}$, which aquires $\hbox{deg}} \def\Pic{\hbox{Pic}} \def\Jac{\hbox{Jac}(Q)\cdot \hbox{deg}} \def\Pic{\hbox{Pic}} \def\Jac{\hbox{Jac}(R)\cdot
\hbox{deg}} \def\Pic{\hbox{Pic}} \def\Jac{\hbox{Jac}(S)\cdot \hbox{deg}} \def\Pic{\hbox{Pic}} \def\Jac{\hbox{Jac}(T) = 4$ ordinary double points. However, these four
ordinary double points all lie on certain rational curves, and after
resolution of singularities, they are components of the $\bf IV^*$ fiber.
Since we
have a constant modulus fibration, fiberwise duality implies that
$A_2^4\oplus E_6$ is
indeed the gauge group of the heterotic string {\it after} the conifold
transition. This kind of information is new and exciting from the point of
view of physics.
\subsection{Elliptic fibrations}
More recently not only K3-fibrations have become of importance
in string theory, but also elliptic fibrations in various dimensions.
In the framework fo F-theory \cite{MV} a more general type of
compactification of the ten-dimensional string (of type IIB) is
considered in which the dilaton field of the string is not
assumed to be constant, as in conventional compactifications,
but instead can vary.
In this solution of the string equations this dilaton is assumed
to take values in an elliptic curve, thus leading to description
of these vacua as 12-dimensional elliptic fibrations.
Interesting theories then can be described by elliptically fibered
Calabi-Yau manifolds of complex dimensions two, three and four.
These new vacua of the type IIB string are conjectured to be dual
to the most general compactification possible for the heterotic
string, based on stable vector bundles with vanishing first
Chern class.
Of particular interest in this context
are Calabi-Yau fourfolds which not only
are elliptic but also K3-fibered. Such spaces lead to a four-dimensional
compactification of F-theories which are expected to be dual
to stable vector bundles over Calabi-Yau threefolds which are elliptic.
Clearly our twist construction provides a systematic method for
building and analyzing such varieties.
In this general framework the question of phase transitions rises
again in the context of the possible connectedness of these
fourdimensional string theory ground states. This problem has
been addressed in \cite{bls97} in the framework of 4D F-theory.
It was shown there that there indeed exists a generalization
of the conifold transition for Calabi-Yau fourfolds. In the
context of these higher dimensional varieties however the
singularities are no longer nodes but the
manifolds degenerate at curves of in general high genus.
A simple example of such a transition between CY$_3$ fibered
fourfolds is provided by \cite{bls97}
\begin{equation}
\fP_{(8,8,4,2,1,1)}[24] \longleftrightarrow} \def\ra{\rightarrow} \def\Ra{\Rightarrow
\matrix{\fP_{(1,1)}\hfill \cr \fP_{(8,8,4,2,1,1)}\cr }
\left[\matrix{1&1\cr 8&16\cr}\right].
\end{equation}
In this transition the singular locus is given by the smooth
curve $\Sigma =\fP_{(4,2,1,1)}[16~~16]$ of genus $g=385$.
At the singular configuration of this space the CY$_3$ fiber
degenerates into a conifold configuration with 32 nodes.
We can obtain this singular fiber by either deforming the
generic hypersurface fiber $\fP_{(4,4,2,1,1)}[12]$ which is
obtained from the twist map
\begin{equation}
\fP_{(2,1,1)}[12]\times \fP_{(2,2,1,1)}[6] ~\longrightarrow} \def\sura{\twoheadrightarrow ~
\fP_{(4,4,2,1,1)}[12]
\end{equation}
i.e. by collapsing 32 three-cycles, or by collapsing 32 two-cycles
of the generic quasismooth codimension two complete intersection
CY which leads to the codim$_{\fC}=2$ fourfold via the
twist map
\begin{equation}
\fP_{(2,1,1)}[8] \times
\matrix{\fP_{(1,1)}\hfill \cr \fP_{(4,4,2,1,1)}\cr}
\left[\matrix{1&1\cr 4&8\cr}\right]
~\longrightarrow} \def\sura{\twoheadrightarrow ~
\matrix{\fP_{(1,1)}\hfill \cr \fP_{(8,8,4,2,1,1)}\cr}
\left[\matrix{1&1\cr 8&16\cr}\right]
\end{equation}
From the analysis above we expect that via our twist map
many of the aspects of the conifold transitions have
important implications for the transitions among Calabi-Yau
fourfolds and, via the conjectured duality between F-theory
and the heterotic string compactified on stable vector bundles
$V~\rightarrow ~$CY$_3$ also for transitions between stable
vector bundles.
\smallskip
\noindent Authors' addresses:
\noindent
{\it Bruce Hunt \\ Max-Planck-Institut f\"ur Mathematik in den
Naturwissenschaften \\
Inselstr. 22-26, D-04103 Leipzig \\ \\ Rolf Schimmrigk \\ Physics
Institute, Bonn University \\ Nu\ss allee 12, D-53115 Bonn \\
and \\ Dept.\ of Physics and Astronomy, Indiana University South Bend
\\ 1700 Mishawaka Av., South Bend, IN 46634, USA }
|
\section{introduction}
In 1972, Kip Thorne proposed the hoop conjecture \cite{hc};
{\it horizons form when and only when a mass $M$ gets compacted into a
region whose circumference in every direction is
${\cal C} \lesssim 4\pi M$.}
However, it took nearly a decade until
the studies
appeared which directly address the conjecture.
This may be because of the vagueness of the statement, i.e.,
the circumference ${\cal C}$ of the horizon and the mass $M$ of a body
are ambiguously defined.
For axisymmetric bodies, the definition of ${\cal C}$ is rather simple
as either the minimum equatorial circumference or the minimum polar
circumference\cite{nst,fla}. Hence there exist a lot of works both by
numerical and by analytical approaches addressing the hoop conjecture in
the axisymmetric system\cite{nst,st,ahst,cnns}.
For non-axisymmetric bodies, on the other hand, we didn't have any
satisfactory definition of ${\cal C}$ until we proposed a
definition using closed geodesics\cite{cnns}.
In this paper we demonstrate how to calculate ${\cal C}$
in numerically generated spacetime without axisymmetry but with a
discrete symmetry and assess the validity of the hoop conjecture.
We study the formation of apparent horizons because apparent horizons
can be identified locally and provide the practical definition
of a black hole. It is important to emphasize here that event horizons
may clothe the singularities even if apparent horizons do not appear
on the spatial slices considered. The existence of an apparent horizon is a
sufficient condition for the formation of an event horizon\cite{he}.
Unfortunately there exists an obstacle to study the formation of apparent
horizons numerically because we do not have an efficient and robust
method to find apparent horizons in general non-axisymmetric spaces,
although there has been recent progress\cite{nko,bcsst,3d,gun}.
This is an important aspect because we are
interested in the existence or nonexistence of an apparent horizon.
On the other hand, if we allow spaces to have a
discrete symmetry (but not continuous one), we can solve the equation
for the surface of an apparent horizon, which is generally a nonlinear
elliptic equation, as a boundary value problem\cite{3d}. Then the
numerical treatment of the problem becomes quite similar to that in
the axisymmetric case.
In this paper, taking advantage of recent progress both in the
definition of the circumference and in finding apparent horizons
numerically,
we study, for the first time, the condition
for the formation of apparent horizons in the light of the hoop conjecture
in spaces without axisymmetry but with a discrete symmetry.
We present
the momentarily static vacuum configurations.
We consider two families of configurations as a
demonstration: black holes distributed on a ring, and black holes
on a spherical surface. We increase the number of black holes keeping
the total mass constant and explore the existence of a common apparent
horizon. As for the definition of a mass we simply adopt the
ADM mass, because of its uniqueness and definiteness at least for the
initial data on a spacelike hypersurface.
This paper is organized as follows.
In Sec.2, first we review the method for finding apparent horizons
in spaces with a discrete symmetry, and then
we also review a definition of the circumference of a
body which has been proposed by us, and
finally we present the results of our
numerical analysis of the initial data. Section 3 is devoted to
summary.
\section{initial data analysis}
\subsection{Finding Apparent Horizons}
A marginally trapped surface is a closed 2-surface $S$ where
the expansion $\Theta $ of future-directed
outgoing null vectors $\ell^{\mu}$ normal to it vanishes\cite{he}.
An apparent horizon is defined as the outer boundary of a connected
component of the trapped region. When the apparent horizon forms, there
always exists the event horizon enclosing it if
there is no naked singularity and the null convergence condition
is satisfied\cite{he}.
Let $s^{\mu}$ be the outward-pointing spacelike unit normal to $S$
and $n^{\mu}$ be the unit normal to a time slice. Then $\ell^{\mu}$
can be written as $\ell^{\mu}=n^{\mu}+s^{\mu}$ and thus
\begin{equation}
\Theta=D_{i}s^i+K_{ij}s^is^j-K=0
\label{expansion}
\end{equation}
on $S$.
In this paper, we consider the conformally flat space
for simplicity
\begin{equation}
dl^2=\psi^4 f_{ij}dx^idx^j,
\end{equation}
where $f_{ij}$ denotes the flat metric. Then Eq.(\ref{expansion})
is rewritten as
\begin{eqnarray}
\Theta=&-&{r\over \psi^2(r^2+r^2_{\theta})^{3/2}}
\Bigl[r_{\theta \theta}+r_{\theta}\cot \theta +{r_{\phi \phi}\over
\sin ^2\theta} -2 r -{3\over r}\left(r_{\theta}^2+{r_{\phi}^2\over
\sin ^2\theta} \right) \nonumber\\
&-& 4\left( {\psi_{r}\over \psi}-{\psi_{\theta}\over \psi}
{r_{\theta}\over r^2}-{\psi_{\phi}\over \psi}{r_{\theta}\over
r^2 \sin ^2\theta}\right)\left(r^2+r_{\theta}^2+{r_{\phi}^2\over
\sin^2 \theta}\right)
+{r_{\theta}^2\over r^2}\left( r_{\theta}\cot \theta + {r_{\phi
\phi}\over \sin ^2 \theta}\right)\nonumber\\
&-&2{r_{\theta }r_{\phi}\over r^2 \sin ^2 \theta}\left(r_{\theta \phi}-
r_{\phi} \cot \theta \right)+{r_{\theta\theta}r_{\phi}^2\over r^2
\sin ^2 \theta}\Bigr]
+K_{ij}s^is^j-K=0,
\label{ah}
\end{eqnarray}
where $r(\theta,\phi)$ parameterizes the surface of the apparent horizon
and $r_{\theta}\equiv \partial r/\partial \theta$. This is the
elliptic partial differential equation about $\theta$ and $\phi$.
It is difficult in general to solve numerically Eq.(\ref{ah})
efficiently and robustly, and
several methods have been proposed so far\cite{nko,bcsst,3d,gun}.
Since we are interested in the criterion of the formation of horizons,
it is important to find horizons robustly. In axisymmetric spaces,
Eq.(\ref{ah}) can be solved as a two-point boundary value problem
\cite{sasaki}. And it is proved to be a robust method to find apparent
horizons since the regularity of
the apparent horizon on the symmetry axis is
automatically guaranteed. It is desirable to
solve Eq.(\ref{ah}) in a similar manner in the non-axisymmetric space.
To do so, we assume that the space has a discrete symmetry such that
$r(\theta,\phi +\pi)=r(\theta,\phi)$. We also assume the space has a
reflection symmetry about $\theta=\pi/2$ plane. The latter
assumption is just for simplicity and is not essential for the method.
Then the axis $\theta=0$ becomes the symmetry axis as in the
axisymmetric case. Thus we can specify the boundary condition at
$\theta=0,\pi/2$ as
\begin{equation}
r_{\theta}=0.
\label{bc}
\end{equation}
Hence we can solve Eq.(\ref{ah}) as a boundary value
problem\cite{3d} although the space is not axisymmetric.
By finite-differencing Eq.(\ref{ah}) with taking account of the discrete
symmetry and the boundary condition Eq.(\ref{bc}), it becomes a matrix
equation, and we solve it iteratively.
The detailed numerical
implementation of the method is given in \cite{3d}.
When the apparent horizon forms, we compute its
area $A$
\begin{equation}
A=4\int_{0}^{\pi}\int_{0}^{\pi/2}\psi^4r^2\sin\theta d\phi d\theta
\sqrt{1+{r_{\theta}^2\over r^2}+{r_{\phi}^2\over r^2\sin^2\theta}}.
\end{equation}
In the numerical results shown below, we discretize $\theta$ and
$\phi$ as follows
\begin{equation}
\theta_{i}=\left(i-{1\over 2}\right){\pi\over 2N_{\theta}},
~~\phi_{j}=\left(j-{1\over 2}\right){\pi \over N_{\phi}},
\end{equation}
with $i=1,\dots, N_{\theta}$, and $j=1,\dots, N_{\phi}$.
We typically take the grid numbers $N_{\theta}=N_{\phi}=64$ and
sometimes $N_{\theta}=N_{\phi}=128$ when a higher resolution is
required.
\subsection{Calculating Hoop}
To gauge the hoop conjecture, we calculate the proper lengths of
geodesics of various orientations that enclose the entire
configuration.
Due to the symmetry such that the reflection symmetry about the
equatorial plane and the discrete symmetry about the $\theta=0$ axis,
the equatorial ($\theta=\pi/2$) plane and some
$\phi=$constant surfaces take the special position. The appropriate
circumference should be the maximum of (i) the minimum equatorial
circumference ${\cal C}^{min}_{eq}$ and (ii) the minimum polar circumference
along $\phi=$ constant plane ${\cal C}^{min}_{pol}$.
The geodesic equation on the equatorial plane is given by
\begin{equation}
r_{\phi\phi}=-2 {\psi_{\phi}\over \psi}{r_{\phi}^3\over r^2}
+\left( 2{\psi_r\over \psi}+{2\over
r}\right)r_{\phi}^2
-2{\psi_{\phi}\over \psi} r_{\phi}+
\left(2{\psi_{r}\over \psi} +{1\over r}\right)r^2.
\label{geo:eq}
\end{equation}
We note that ${\cal C}^{min}_{eq}$ is well-defined only if the equatorial
plane symmetry is assumed as in the recent 3D dynamical calculations
on coalescing binary NSs/BHs\cite{binary}.
On the other hand, geodesics along the $\phi=$ constant plane
do not exist in general. However, if we restrict ourselves to
the plane with the symmetry such that $\psi_{\phi}=0$, then such geodesics do
exist and follow the equation:
\begin{equation}
r_{\theta\theta}=-2 {\psi_{\theta}\over \psi}{r_{\theta}^3\over r^2}
+\left( 2{\psi_r\over \psi}+{2\over
r}\right)r_{\theta}^2
-2{\psi_{\theta}\over \psi} r_{\theta}+
\left(2{\psi_{r}\over \psi} +{1\over r}\right)r^2.
\label{geo:pol}
\end{equation}
Eq.(\ref{geo:eq}) and
Eq.(\ref{geo:pol}) are the same
except for the replacement $\phi \rightarrow \theta$,
and we solve them by the fourth-order Runge-Kutta method.
We vary $r$ and $r_{\phi}(r_{\theta})$ at $\phi=0(\theta=0)$
and compute ${\cal C}$ encompassing all black holes using
Eq.(\ref{geo:eq})(Eq.(\ref{geo:pol})) up to
$\phi=\pi(\theta=\pi/2)$. In this way we obtain the minimum
circumference:
\begin{eqnarray}
{\cal C}^{min}_{eq}&=&2\int^{\pi}_0\psi^2\left(r^2+r_{\phi}^2\right)^{{1\over
2}}d\phi,\\
{\cal C}^{min}_{pol}&=&4 \int^{\pi/2}_0\psi^2\left(r^2+
r_{\theta}^2\right)^{{1\over
2}}d\theta.
\end{eqnarray}
We note that when we compute the minimum
polar circumference, we first minimize the circumference with being $\phi$
fixed and then maximize over $\phi$ which satisfies $\psi_{\phi}=0$.
\subsection{Numerical Results}
Now we construct the vacuum initial data. We consider the
time-symmetric initial data, that is, $K_{ij}=0$.
Then the only constraint equation we have to solve is the Hamiltonian
constraint equation
\begin{equation}
\Delta \psi=0,
\label{hamilt}
\end{equation}
where $\Delta$ denotes the flat Laplacian. We consider the following
configurations: (1) black holes distributed on a ring, (2) black holes
on a spherical surface.
\subsubsection{Case (1)}
We show the numerical results for the case (1): black holes
distributed on a ring.
We consider $N$ black holes (where $N$ is an even number) of an equal mass
distributed on a ring of radius $a$ such that each black hole's
Cartesian coordinate is $(a\cos(2\pi l/N),a\sin(2\pi l/N),0)$,
where $l=0,\dots, N-1$. Then the solution of the
constraint Eq.(\ref{hamilt}) is
\begin{equation}
\psi(r,\theta,\phi)=
1+ {M\over 2N}\sum_{l=0}^{N-1}{1\over \sqrt{r^2 -2ar \sin\theta
\cos\left(\phi-2\pi l/N\right) +a^2}},
\end{equation}
where $M$ is the ADM mass of the total system. In the limit of large
$N$ the configuration becomes the singular ring, and the solution
becomes
\begin{equation}
\psi(r,\theta)=1+ {M\over \pi\sqrt{r^2+2ar\sin\theta +a^2}}K(\kappa),
\end{equation}
where $K(\kappa)$ is the elliptic integral of the first kind with
\begin{equation}
\kappa^2={4ar\sin\theta\over r^2+2ar\sin\theta +a^2}.
\end{equation}
The results are shown in Fig.1 and Table 1.
The top view of
the apparent horizon as well as the bird's-eye view of it is shown.
The shape of the apparent horizon indeed respects the discrete
symmetry of the configuration.
As $N$ increases, the shape becomes pancake-like.
The maximum of $({\cal C}^{min}_{eq},{\cal C}^{min}_{pol})$ is shown in the
table. We calculate the circumference ${{\cal C}}$ of the system irrespective
of the existence of an apparent horizon. Even if the apparent horizon
forms, the appropriate circumference is generally not located on it.
The circumference on the apparent horizon can be significantly
larger\cite{nst,cnns,bs}.
When the apparent horizon forms, its area $A$ normalized by
$16\pi M^2$
is computed because in the time-symmetric initial data the following
inequality holds if the final state is stationary:
\begin{equation}
A \leq A_{EH} \leq A_f(=16\pi M_f^2) \leq 16\pi M^2.
\end{equation}
Here $A_{EH}$ is the area of the event horizon. We note that the
existence of an apparent horizon implies the existence of the event
horizon enclosing it\cite{he}. From this and the fact that \
the apparent horizon is the minimum surface in the time-symmetric
initial data\cite{gibbons}, the first inequality holds.
$A_f$ is the final stationary state black hole's area, and the second
inequality follows from the area theorem\cite{he}. From the
uniqueness theorem\cite{israel}, the final state is the
Schwarzschild black hole with mass $M_f$ because we are considering
the non-rotating system. $M_f$ should not be larger than the initial mass
$M$ because gravitational radiation conveys the energy during the
dynamical evolution.
In the cases $N=2$ and $N
\rightarrow \infty$, we also show the numerical results using a
2D(axisymmetric) apparent horizon finder\cite{sasaki}
for comparison and the check of our numerical results. We find good
agreement.
\subsubsection{Case (2)}
Next we show the numerical results for the case (2): black holes
on a spherical surface.
Black holes are distributed on a sphere of radius $a$ such that each
black hole's Cartesian coordinate is
\begin{equation}
(a\sin(\pi k/n)\cos(2\pi l/n),a\sin(\pi k/n)
\sin(2\pi l/n),a\cos(\pi k/n)),
\end{equation}
where $k=1,\dots,n-1,l=0,\dots, n-1$. We also consider two additional
black holes located at $\theta=0$, and $\pi$ on the surface
so that in the large $N$ limit the configuration becomes a singular
spherical surface. The total number of black holes
is then $N=n(n-1)+2$. We assume $n$ is even number. The solution of the
constraint Eq.(\ref{hamilt}) is simply
\begin{eqnarray}
\psi(r,\theta,\phi)&=&1+ {M\over 2N}\left({1\over
\sqrt{r^2+2ar\cos\theta+a^2}}+{1\over
\sqrt{r^2-2ar\cos\theta+a^2}}
\right)\nonumber\\
&+&{M\over 2N}\sum_{k=1}^{n-1}\sum_{l=0}^{n-1}{1\over
\sqrt{r^2+a^2-2ar \left( \sin\theta\sin(\pi k/n)\cos(\phi-2\pi l/n)
+\cos\theta\cos(\pi k/n)\right)}}.
\end{eqnarray}
The results are shown in Fig.2 and Table 2. For smaller $N$
the apparent horizon is bumpy.
Since the system becomes spherically symmetric in the large $N$ limit,
$A/16\pi M^2$ becomes close to 1 as $N$ increases and ${\cal C}$ does
not clearly distinguish the existence or non-existence of the apparent
horizon.
\section{summary}
We tested the hoop conjecture in spaces without axisymmetry but with a
discrete symmetry. The existence or nonexistence of an apparent
horizon encompassing all black holes is qualitatively consistent with
the hoop conjecture proposed by Thorne. From Table 1 and Table 2, we
find that if the circumference ${\cal C}$ satisfies
${\cal C}/4\pi M \lesssim 1.168$, then the configuration will be
surrounded by a common apparent horizon.
In non-axisymmetric spaces with a discrete symmetry,
searching for hoops
is less computationally demanding than searching for apparent
horizons: just solving the ordinary differential equation.
Further if the equatorial plane symmetry is assumed
as in the recent 3D dynamical calculations on
coalescing binary NSs/BHs\cite{binary},
${\cal C}_{eq}^{min}$ is then well-defined, and therefore even in
more general spaces the hoop concept can be a useful diagnostic for
the final fate of the collapsed object.
\acknowledgments
The author would like to thank Professor T.Nakamura for encouragement
and Professor N.Sugiyama for a careful reading of the manuscript.
He is also grateful to an anonymous referee for useful comments which
improved the manuscript.
This work was supported in part by YITP and in part by a JSPS
Fellowship for Young Scientists under grant No.3596.
|
\section{Introduction}
Inspiraling neutron star binaries are expected to be among the
strongest sources of gravitational radiation that could be detected by
the interferometric detectors currently under construction (GEO600,
LIGO, TAMA and Virgo). These binary systems are therefore subject to
numerous theoretical studies. Among them are
(i) Post-Newtonian (PN) analytical treatments (e.g. \cite{BlancI98},
\cite{BuonaD99}, \cite{Tanig99}) and (ii)
fully relativistic
hydrodynamical treatments, pioneered by the works of Oohara and
Nakamura (see e.g. \cite{OoharN97}) and Wilson et al.
\cite{WilsoM95,WilsoMM96}. The most recent numerical calculations,
those of Baumgarte et al. \cite{BaumgCSST97,BaumgCSST98a} and
Marronetti et al. \cite{MarroMW98}, rely on the approximations of (i)
a quasiequilibrium state and (ii) of synchronized binaries. Whereas
the first approximation is well justified before the innermost stable
orbit, the second one does not correspond to physical situations, since
it has been shown that the gravitational-radiation driven evolution is
too rapid for the viscous forces to synchronize the spin of each
neutron star with the orbit \cite{Kocha92,BildsC92} as they do for
ordinary stellar binaries. Rather, the viscosity is negligible and
the fluid velocity circulation (with respect to some inertial
frame) is conserved in these systems. Provided that the initial spins
are not in the millisecond regime, this means that close binary
configurations are better approximated by zero vorticity (i.e. {\em
irrotational}) states than by synchronized states.
Dynamical calculations by Wilson et al.
\cite{WilsoM95,WilsoMM96} indicate that the neutron stars may
individually collapse into a black hole prior to merger. This
unexpected result has been called into question by a number of authors
(see Ref.~\cite{MatheMW98} for a summary of all the criticisms and
their answers). Recently Flanagan \cite{Flana99} has found an error in
the analytical formulation used by Wilson et al. \cite{WilsoM95,WilsoMM96}.
This error may be responsible of the observed radial instability.
As argued by Mathews et al.~\cite{MatheMW98}, one way
to settle this crucial point is to perform computations of relativistic
irrotational configurations. We have performed recently
such computations \cite{BonazGM99}. They show no compression of the stars,
although the central density decreases much less than in the
corotating case. In the present report, we give more details about the
results presented in Ref.~\cite{BonazGM99} and extend them to the
compactification ratio $M/R=0.17$ (the results of Ref.~\cite{BonazGM99}
have been obtained for a compactification ratio $M/R=0.14$).
\section{Analytical formulation of the problem}
\subsection{Basic assumptions} \label{s:assumpt}
We have proposed a relativistic formulation for quasiequilibrium
irrotational binaries \cite{BonazGM97b} as a generalization of the
Newtonian formulation presented in Ref.~\cite{BonazGHM92}. The method was
based on one aspect of irrotational motion, namely the {\em
counter-rotation} (as measured in the co-orbiting frame) of the fluid
with respect to the orbital motion (see also Ref.~\cite{Asada98}).
Since then, Teukolsky~\cite{Teuko98} and Shibata~\cite{Shiba98} gave two
formulations based on the definition of irrotationality, which implies
that the specific enthalpy times the fluid 4-velocity is the gradient
of some scalar field \cite{LandaL89} ({\em potential flow}). The three
formulations are equivalent; however the one given by Teukolsky and by
Shibata greatly simplifies the problem. Consequently we used it in the present
work.
The irrotational hypothesis amounts to say that the co-momentum density is
the gradient of a potential
\begin{equation} \label{e:irrot}
h \, {\bf u} = \nabla \Psi \ ,
\end{equation}
where $h$ and $\bf u$ are respectively the fluid specific enthalpy and
fluid 4-velocity.
\begin{figure}
\centering
\epsfig{figure=07_SBONAZZOLA1_1.eps,height=6cm}
\caption{Spacetime foliation $\Sigma_t$, helicoidal Killing vector
$l^\alpha$ and its trajectories $x^{i'} = {\rm const}$, which are the
worldlines of the co-orbiting observer (4-velocity: $v^\alpha$).
Also shown are the rotating-coordinate shift vector $B^\alpha$
and the unit future-directed vector $n^\alpha$, normal to the spacelike
hypersurface $\Sigma_t$ (Figure from Ref.~\cite{BonazGM97b}).}\label{f:helic}
\end{figure}
Beside the physical assumption (\ref{e:irrot}), two simplifying approximations
are introduced:
\begin{enumerate}
\item The spacetime is supposed to have a helicoidal symmetry
\cite{BonazGM97b}, which means
that the orbits are exactly circular and that the gravitational radiation
content of spacetime is neglected.
\item The spatial 3-metric is assumed to be conformally flat (Wilson-Mathews
approximation \cite{WilsoM89,WilsoM95}), so that the full spacetime
metric reads
\begin{equation} \label{e:met}
ds^2 = -(N^2 - B_i B^i) dt^2 - 2 B_i dt\, dx^i
+ A^2 f_{ij} dx^i\, dx^j \ ,
\end{equation}
where $f_{ij}$ is the flat space metric.
\end{enumerate}
The Killing vector corresponding to hypothesis (i) is
denoted by $\bf l$ (cf. Fig.~\ref{f:helic}).
Approximation (i) is physically well motivated, at least up to the
innermost stable orbit. Regarding the second approximation, it is to be
noticed that (i) the 1-Post Newtonian (PN) approximation
to Einstein equations fits this form, (ii) it is exact for arbitrary
relativistic spherical configurations and (iii) it is very accurate for
axisymmetric rotating neutron stars \cite{CookST96}. An interesting discussion
about some justifications of the Wilson-Mathews approximation may be
found in \cite{MatheMW98}. A stronger justification may be obtained by
considering the 2.5-PN metric obtained by Blanchet et al.
\cite{BlancFP98} for point mass binaries. Using Eq.~(7.6) of
Ref.~\cite{BlancFP98}, the deviation from a conformally flat 3-metric
(which occurs at the 2-PN order) can be computed at the location of one
point mass (i.e. where it is expected to be maximal),
the 3-metric $h_{ij}$ being written as
\begin{equation} \label{e:met-PN}
h_{ij} = A^2 \, f_{ij} + h_{ij}^{\rm 2PN} + h_{ij}^{\rm 2.5PN} \ .
\end{equation}
The result is shown in Table~\ref{t:PN} for two stars of $1.4 M_\odot$ each.
It appears that at a separation as close as $30{\ \rm km}$, where the two
stars certainly almost touch each other, the relative deviation from a
conformally flat 3-metric is below $2\%$.
\begin{table}[t]
\begin{center}
\begin{tabular}{*{6}{c}}
\hline
\\[0.5ex]
$d$ & $v/c$ & $\omega_{\rm 2PN}$ & Conformal fact.
& $\left| h_{ij}^{\rm 2PN} \right| /A^2$
& $\left| h_{ij}^{\rm 2.5PN} \right| /A^2$ \\
$\rm [km]$ & & [rad/s] & $A^2$ & & \\[0.5ex]
\hline
\\[0.5ex]
100 & 0.10 & 579 & 1.04 & $1.8\times 10^{-3}$ & $6.3\times 10^{-4}$
\\[0.5ex]
50 & 0.13 & 1572 & 1.09 & $6.8\times 10^{-3}$ & $3.3\times 10^{-3}$
\\[0.5ex]
40 & 0.14 & 2166 & 1.11 & $1.0\times 10^{-2}$ & $5.6\times 10^{-3}$
\\[0.5ex]
30 & 0.16 & 3387 & 1.15 & $1.8\times 10^{-2}$ & $1.1\times 10^{-2}$
\\[0.5ex]
\hline
\end{tabular}
\vspace{3mm}
\caption{Deviation from a conformally flat 3-metric at 2-PN at 2.5-PN order
for point mass binaries of $M=1.4\, M_\odot$ each. $d$ is the coordinate
separation between the two stars, $v$ is the coordinate orbital velocity
of one star, $\omega_{\rm 2PN}$ is the orbital angular velocity, the other
notations are defined by Eq.~(\ref{e:met-PN}). The metric at the
2.5PN level is taken from Blanchet et al.~\cite{BlancFP98} and is computed
at the location of the point masses.}
\label{t:PN}
\end{center}
\end{table}
\subsection{Equations to be solved}
We refer to Ref.~\cite{BonazGM99} for the presentation of the partial
differential equations (PDEs) which result from the assumptions presented above.
In the present report, let us simply mention a point which seems to have
been missed by various authors: the existence of the first integral of motion
\begin{equation} \label{e:int_prem}
h \, {\bf l} \cdot {\bf u} = {\rm const.}
\end{equation}
does not result solely from the existence of the helicoidal
Killing vector $\bf l$. Indeed, Eq.~(\ref{e:int_prem})
is not merely the relativistic generalization of the Bernoulli
theorem which states that $h \, {\bf l} \cdot {\bf u}$ is constant along
each single streamline and which results directly from the existence of
a Killing vector without any hypothesis on the flow. In order for the
constant to be uniform over the streamlines (i.e. to be a constant over
spacetime), as in Eq.~(\ref{e:int_prem}), some additional property of the
flow must be required. One well known possibility is {\em rigidity} (i.e.
$\bf u$ colinear to $\bf l$) \cite{Boyer65}. The alternative property
with which we are concerned here is {\em irrotationality}
[Eq.~(\ref{e:irrot})]. This was first pointed out by Carter
\cite{Carte79}.
\begin{figure}
\centering
\epsfig{figure=07_SBONAZZOLA1_2.eps,height=7cm}
\caption{Logarithm of the relative global error of the numerical solution
with respect to the number of degrees of freedom in
$\theta$ for a Roche ellipsoid for an equal mass binary system and
$\Omega^2/(\pi G\rho) = 0.1147$ (the numbers of degrees of freedom in the
other directions are $N_r = 2N_\theta-1$ and $N_\varphi = N_\theta -1$).
Also shown is the error in the
verification of the virial theorem (Figure from Ref.~\cite{BonazGM98a}).}
\label{f:test:roche}
\end{figure}
\section{Numerical procedure}
\subsection{Description}
The numerical procedure to solve the PDE system
is based on the multi-domain spectral method presented in Ref.~\cite{BonazGM98a}.
We simply recall here some basic features of the method:
\begin{itemize}
\item Spherical-type coordinates $(\xi,\theta',\varphi')$ centered on each
star are used: this ensures
a much better description of the stars than with Cartesian coordinates.
\item These spherical-type coordinates are surface-fitted coordinates: i.e.
the surface of each star lies at a constant value of the coordinate
$\xi$ thanks to a mapping $(\xi,\theta',\varphi')\mapsto (r,\theta,\varphi)$
(see \cite{BonazGM98a} for details about this mapping). This
ensures that the spectral method applied in each domain is free from any
Gibbs phenomenon.
\item The outermost domain extends up to spatial infinity, thanks to the
mapping $1/r = (1-\xi)/(2R_0)$. This enables to put exact boundary conditions
on the elliptic equations for the metric coefficients: spatial infinity
is the only location where the metric is known in advance (Minkowski metric).
\item Thanks to the use of a spectral method \cite{BonazGM99b}
in each domain, the numerical
error is {\em evanescent}, i.e. it decreases exponentially with the number
of coefficients (or equivalently grid points) used in the spectral expansions,
as shown in Fig.~\ref{f:test:roche}.
\end{itemize}
The PDE system to be solved being non-linear, we use an iterative
procedure. This procedure is sketched in Fig.~\ref{f:iter}.
The iteration is stopped when the relative difference in the enthalpy field
between two successive steps goes below a certain threshold, typically
$10^{-7}$ (cf. Fig.~\ref{f:converge}).
\begin{figure}
\centering
\epsfig{figure=07_SBONAZZOLA1_3.eps,height=8cm}
\caption{Schematic representation of the iterative procedure used in the
numerical code.}
\label{f:iter}
\end{figure}
\begin{figure}
\centering
\epsfig{figure=07_SBONAZZOLA1_4.eps,height=7cm}
\caption{Convergence (measured by the relative difference in the enthalpy field
between two successive steps) of the iterative procedure for one of the
irrotational models considered in Ref.~\cite{BonazGM99}. The bump around
the 90th step corresponds to the switch on of the procedure of convergence
toward a given baryon mass.}
\label{f:converge}
\end{figure}
The numerical code is written in the {\sc LORENE} language
\cite{MarckG99}, which is a C++ based language for numerical relativity.
A typical run make use of $N_r = 33$, $N_\theta = 21$, and
$N_\varphi=20$ coefficients (= number of
collocation points, which may be seen as number of grid points) in
each of the domains on the multi-domain spectral method. 8 domains are
used : 3 for each star and 2 domains centered on the intersection between
the rotation axis and the orbital plane. The corresponding memory requirement
is 155 MB. A computation involves $\sim 250$ steps (cf. Fig.~\ref{f:converge}),
which takes 9~h~30~min on one CPU of a SGI Origin200 computer
(MIPS~R10000 processor at 180 MHz).
Note that due to the rather small memory requirement,
runs can be performed in parallel on a multi-processor platform.
This especially useful to compute sequences of configurations.
\subsection{Tests passed by the code}
In the Newtonian and incompressible
limit, the analytical solution constituted by a Roche ellipsoid is
recovered with a relative accuracy of $\sim 10^{-9}$, as shown in
Fig.~\ref{f:test:roche}. For compressible and irrotational Newtonian
binaries, no analytical solution is available, but the virial theorem
can be used to get an estimation of the numerical error: we found that the
virial theorem is satisfied with a relative accuracy of $10^{-7}$. A
detailed comparison with the irrotational Newtonian configurations
recently computed by Uryu \& Eriguchi \cite{UryuE98a,UryuE98b} will be
presented elsewhere. Regarding the relativistic case, we
have checked our procedure of resolution of the gravitational field
equations by comparison with the results of Baumgarte et
al.~\cite{BaumgCSST98a} which deal with corotating binaries [our code
can compute corotating configurations by setting to zero the
velocity field of the fluid with respect to the co-orbiting observer].
We have performed the comparison with
the configuration $z_A=0.20$ in Table~V of Ref.~\cite{BaumgCSST98a}. We
used the same equation of state (EOS) (polytrope with $\gamma=2$), same
value of the separation $r_C$ and same value of the maximum density parameter
$q^{\rm max}$. We found a relative discrepancy of $1.1\%$ on $\Omega$,
$1.4\%$ on $M_0$, $1.1\%$ on $M$, $2.3\%$ on $J$, $0.8\%$ on $z_A$,
$0.4\%$ on $r_A$ and $0.07\%$ on $r_B$ (using the notations of
Ref.~\cite{BaumgCSST98a}).
\section{Numerical results} \label{s:num_res}
\subsection{Equation of state and compactification ratio}
As a model for nuclear matter,
we consider a polytropic equation of state (EOS) with an adiabatic index
$\gamma=2$:
\begin{equation} \label{e:eos}
p=\kappa (m_{\rm B} n)^\gamma \ , \qquad
e=m_{\rm B} n + p/(\gamma-1) \ ,
\end{equation}
where $p$, $n$, $e$ are respectively the fluid pressure, baryon density and
proper energy density, and $m_{\rm B} = 1.66\times 10^{-27} {\rm\ kg}$,
$\kappa = 1.8 \times 10^{-2} {\ \rm J\, m}^3{\rm kg}^{-2}$. This EOS is the
same as that used by Mathews, Marronetti and Wilson
(Sect.~IV~A of Ref~\cite{MatheMW98}).
\begin{figure}
\centering
\epsfig{figure=07_SBONAZZOLA1_5.eps,height=7cm}
\caption{Mass as a function of the central energy density for
static isolated neutron stars constructed with the EOS (\ref{e:eos}).
The heavy dots are configurations considered by our group
\cite{BonazGM99} and
Marronetti, Mathews \& Wilson group \cite{MatheMW98}, \cite{MarroMW99}
(see text)
($\rho_{\rm nuc} := 1.66\times 10^{17} {\rm \ kg\, m}^{-3}$).}
\label{f:m_sher}
\end{figure}
The mass -- central density curve of static
configurations in isolation constructed upon this EOS is represented
in Fig.~\ref{f:m_sher}. The three points on this curve corresponds to
three configurations studied by our group and that of Marronetti, Mathews
and Wilson:
\begin{itemize}
\item The configuration of baryon mass $M_{\rm B} = 1.625 \, M_\odot$
and compactification ratio $M/R = 0.14$ is
that considered in the dynamical study of Mathews, Marronetti and Wilson
\cite{MatheMW98} and in the quasiequilibrium study of our group
(Ref.~\cite{BonazGM99} and this paper).
\item The configuration of baryon mass $M_{\rm B} = 1.85 \, M_\odot$
and compactification ratio $M/R = 0.17$ is studied in the present paper.
\item The configuration of baryon mass $M_{\rm B} = 1.95 \, M_\odot$
and compactification ratio $M/R = 0.19$ has been studied recently by
Marronetti, Mathews and Wilson \cite{MarroMW99}\footnote{Marronetti et al.
\cite{MarroMW99} use a different value for the EOS constant $\kappa$:
their baryon mass $M_{\rm B} = 1.55 \, M_\odot$ must be rescaled to our
value of $\kappa$ in order to get $M_{\rm B} = 1.95 \, M_\odot$. Apart from
this scaling, this is the same configuration, i.e. it has the same
compactification ratio $M/R = 0.19$ and its relative distance with respect to
the maximum mass configuration, as shown in Fig.~\ref{f:m_sher}, is the same.}
by means of a new code for quasiequilibrium irrotational configurations.
\end{itemize}
\subsection{Irrotational sequence with $M/R=0.14$}
In this section, we give some details about the irrotational sequence
$M_{\rm B} = 1.625 \, M_\odot$ presented in Ref.~\cite{BonazGM99}.
This sequence starts at the coordinate separation $d=110 {\rm\ km}$ (orbital
frequency $f=82{\rm\ Hz}$), where the two stars are almost spherical, and
ends at $d=41 {\rm\ km}$ ($f=332{\rm\ Hz}$), where a cusp appears on the
surface of the stars, which means that the stars start to break.
The shape of the surface at this last point is shown in Fig.~\ref{f:surf-3D}.
\begin{figure}
\centering
\epsfig{figure=07_SBONAZZOLA1_6.eps,height=7cm}
\caption{Shape of irrotational binary neutron stars of baryon mass
$M_{\rm B} = 1.625 \, M_\odot$, when the coordinate separation between
their centers (density maxima) is $41{\rm\ km}$. Only one half of each star
is represented (the part which is above the orbital plane).
The drawing is that of the numerical grid, which coincides with the surface
of the star, thanks to the use of surface-fitted spherical coordinates.}
\label{f:surf-3D}
\end{figure}
The velocity field with respect to the co-orbiting observer, as defined
by Eq.~(52) of Ref.~\cite{BonazGM97b}, is shown in Fig.~\ref{f:vit}.
Note that this field is tangent to the surface of the star, as it must be.
The lapse function $N$ (cf. Eq.~\ref{e:met}) is represented in
Fig.~\ref{f:lapse}. The coordinate system $(x,y,z)$ is centered on
the intersection between the rotation axis and the orbital plane.
The $x$ axis joins the two stellar centers, and the orbital is the $z=0$
plane. The value of $N$ at the center of each star is
$N_{\rm c}=0.64$.
The conformal factor $A^2$ of the 3-metric [cf. Eq.~(\ref{e:met})] is
represented in Fig.~\ref{f:acar}. Its value at the center of each star is
$A^2_{\rm c}=2.20$.
The shift vector of nonrotating coordinates, $\bf N$, (defined by Eq.~(9) of
Ref.~\cite{BonazGM99}) is shown in Fig.~\ref{f:shift}.
Its maximum value is $0.10\, c$.
The $K_{xy}$ component of the extrinsic curvature tensor of the hypersurfaces
$t={\rm const}$ is shown in Fig.~\ref{f:kxy}. We chose to represent the
$K_{xy}$ component because it is the only component of $K_{ij}$ for which
none of the sections in the three planes $x=0$, $y=0$ and $z=0$
vanishes.
\begin{figure}
\centering
\epsfig{figure=07_SBONAZZOLA1_7.eps,height=7cm}
\caption{Velocity field with respect to the co-orbiting observer, for the
configuration shown in Fig.~\ref{f:surf-3D}. The plane of the figure is the
orbital plane. The heavy line denotes the surface of the star.
The companion is located at $x=+41{\rm\ km}$.}
\label{f:vit}
\end{figure}
\begin{figure}
\centering
\epsfig{figure=07_SBONAZZOLA1_8.eps,height=10cm}
\caption{Isocontour of the lapse function $N$ for the
configuration shown in Fig.~\ref{f:surf-3D}. The plots are cross section
is the $x=0$, $y=0$ and $z=0$ planes (note that the $x$ coordinate is
shifted by $20.5{\rm\ km}$ with respect to that of Fig.~\ref{f:vit}).
The dot-dashed line denotes the
boundary between the inner numerical grid and the outer compactified one
(which extends to spatial infinity), for the grid system centered on
the intersection between the rotation axis and the orbital plane.}
\label{f:lapse}
\end{figure}
\begin{figure}
\centering
\epsfig{figure=07_SBONAZZOLA1_9.eps,height=10cm}
\caption{Same as Fig.~\ref{f:lapse} but for the conformal factor $A^2$
of the spatial metric.}
\label{f:acar}
\end{figure}
\begin{figure}
\centering
\epsfig{figure=07_SBONAZZOLA1_10.eps,height=8cm}
\caption{Shift vector of nonrotating coordinates in the orbital plane,
for the configuration shown in Figs.~\ref{f:surf-3D}-\ref{f:acar}.}
\label{f:shift}
\end{figure}
\begin{figure}
\centering
\epsfig{figure=07_SBONAZZOLA1_11.eps,height=10cm}
\caption{Same as Fig.~\ref{f:lapse} but for the component $K_{xy}$ of
the extrinsic curvature tensor. The solid (resp. dashed) lines corresponds
to positive (resp. negative) values of $K_{xy}$.}
\label{f:kxy}
\end{figure}
The variation of the central density along the $M_{\rm B}=1.625\, M_\odot$
sequence is shown in Fig.~\ref{f:cent_dens_1.625}. We have also
computed a corotating sequence for comparison (dashed line in
Fig.~\ref{f:cent_dens_1.625}). In the corotating case, the central density
decreases quite substantially as the two stars approach each other.
This is in agreement with the results of Baumgarte et
al.~\cite{BaumgCSST97,BaumgCSST98a}. In the irrotational case (solid
line in Fig.~\ref{f:cent_dens_1.625}), the central density remains rather
constant (with a slight increase, below $0.1\%$)
before decreasing. We can thus conclude that no tendency to individual
gravitational collapse is found in this case.
This contrasts with results of dynamical calculations by
Mathews et al.~\cite{MatheMW98} which show a central density increase
of $14\%$ for the same compactification ratio $M/R=0.14$.
\begin{figure}
\centering
\epsfig{figure=07_SBONAZZOLA1_12.eps,height=7cm}
\caption{Relative variation of the central energy density $e_{\rm c}$ with
respect to its value at infinite separation $e_{\rm c}^{\rm inf}$ as a
function of the coordinate separation $d$ (or of the orbital frequency
$\Omega/(2\pi)$) for constant baryon mass $M_{\rm B} = 1.625 \,
M_\odot$ sequences. The solid (resp. dashed) line corresponds to a
irrotational (resp. corotating) sequence of coalescing neutron star
binaries (Figure from Ref.~\cite{BonazGM99}).}
\label{f:cent_dens_1.625}
\end{figure}
\subsection{Irrotational sequence with $M/R=0.17$}
In order to investigate how the above result depends on the compactness
of the stars, we have computed an irrotational sequence with a baryon
mass $M_{\rm B} = 1.85\, M_\odot$, which corresponds to a compactification
ratio $M/R=0.17$ for stars at infinite separation
(second heavy dot in Fig.~\ref{f:m_sher}).
The result is compared
with that of $M/R=0.14$ in Fig.~\ref{f:cent_dens_1.625_1.85}.
A very small density increase (at most $0.3\%$)
is observed before the decrease. Note that this density increase remains
within the expected error ($\sim 2\%$, cf. Sect.~\ref{s:assumpt}) induced
by the conformally flat approximation for the 3-metric, so that it cannot
be asserted that this effect would remain in a complete calculation.
Marronetti, Mathews and Wilson~\cite{MarroMW99} have recently computed
quasiequilibrium irrotational configurations by means of a new code.
They use a higher
compactification ratio, $M/R=0.19$ (third heavy dot in Fig.~\ref{f:m_sher}).
They found a central density increase as the orbit shrinks much
pronounced than that we found for the compactification ratio $M/R=0.17$:
$3.5\%$ against $0.3\%$. We will present irrotational sequences with
the compactification ratio $M/R=0.19$ and compare with the results by
Marronetti et al.~\cite{MarroMW99} in a future article.
\begin{figure}
\centering
\epsfig{figure=07_SBONAZZOLA1_13.eps,height=7cm}
\caption{Relative variation of the central energy density $e_{\rm c}$ with
respect to its value at infinite separation $e_{\rm c}^{\rm inf}$ as a
function of the coordinate separation $d$ for constant baryon mass
sequences with
$M_{\rm B} = 1.625 \, M_\odot$ (solid line, same as in
Fig.~\ref{f:cent_dens_1.625}) and
$M_{\rm B} = 1.85\, M_\odot$ (dashed line).}
\label{f:cent_dens_1.625_1.85}
\end{figure}
\section{Innermost stable circular orbit}
An important parameter for the detection of a gravitational wave signal from
coalescing binaries is the location of the innermost stable circular
orbit (ISCO), if any. In Table~\ref{t:isco}, we recall what is known about
the existence of an ISCO for extended fluid bodies. The case of two point
masses is discussed in details in Ref.~\cite{BuonaD99}.
\begin{table}[t]
\begin{center}
\begin{tabular}{*{3}{c}}
\hline
\\[0.5ex]
Model & Existence of an ISCO & References \\[0.5ex]
\hline
\\[0.5ex]
Newtonian corotating & ISCO $\Leftrightarrow \, \gamma > 2$ & \cite{ShibaTN97}
\\[0.5ex]
Newtonian irrotational & ISCO $\Leftrightarrow \, \gamma > 2.4$ &
\cite{UryuE98a}
\\[0.5ex]
GR corotating & ISCO $\Leftrightarrow \, \gamma > 5/3$ &
\cite{BaumgCSST97}, \cite{ShibaTN97}
\\[0.5ex]
GR irrotational & ISCO exists for $\gamma = \infty$ &
\cite{Tanig99}
\\[0.5ex]
\hline
\end{tabular}
\vspace{3mm}
\caption{Known results about the existence of an ISCO for extended fluid
bodies, in terms of the adiabatic index $\gamma$.}
\label{t:isco}
\end{center}
\end{table}
For Newtonian binaries, it has been shown \cite{LaiRS93} that the ISCO
is located at a minimum of the total energy, as well as total angular momentum,
along a sequence at constant baryon number and constant circulation
(irrotational sequences are such sequences). The instability found in this
way is dynamical. For corotating sequences, it is secular instead
\cite{LaiRS93,LaiRS94}.
This turning point method also holds for locating ISCO in
relativistic corotating binaries \cite{BaumgCSST98b}.
For relativistic irrotational configurations, no rigorous theorem has been
proven yet about the localization of the ISCO by a turning point method.
All what can be said is that no turning point is present in the
irrotational sequences considered in Sect.~\ref{s:num_res}:
Fig.~\ref{f:ADM_1.625} shows the variation as the orbit shrinks
of the ADM mass of the spatial hypersurface $t={\rm const}$ (which is a
measure of the total energy, or equivalently of the
binding energy, of the system) for the $M_{\rm B}=1.625\, M_\odot$
sequence. Clearly, the ADM mass decreases
monoticaly, without showing any turning point. Figure~\ref{f:J_1.625}
shows the evolution of the total angular momentum along the same sequence.
Again there is no turning point.
The same behaviour holds for the $M_{\rm B}=1.85\, M_\odot$ sequence, as
shown in Figs.~\ref{f:ADM_1.85} and \ref{f:J_1.85}.
\begin{figure}
\centering
\epsfig{figure=07_SBONAZZOLA1_14.eps,height=6.5cm}
\caption{Half of the ADM mass of the binary system as a function of the
coordinate distance $d$, along the evolutionnary sequence
$M_{\rm B} = 1.625 \, M_\odot$.}
\label{f:ADM_1.625}
\end{figure}
\begin{figure}
\centering
\epsfig{figure=07_SBONAZZOLA1_15.eps,height=6.5cm}
\caption{Total angular momentum of the binary system as a function of the
coordinate distance $d$, along the evolutionary sequence
$M_{\rm B} = 1.625 \, M_\odot$.}
\label{f:J_1.625}
\end{figure}
\begin{figure}
\centering
\epsfig{figure=07_SBONAZZOLA1_16.eps,height=6.5cm}
\caption{Half of the ADM mass of the binary system as a function of the
coordinate distance $d$, along the evolutionary sequence
$M_{\rm B} = 1.85 \, M_\odot$.}
\label{f:ADM_1.85}
\end{figure}
\begin{figure}
\centering
\epsfig{figure=07_SBONAZZOLA1_17.eps,height=6.5cm}
\caption{Total angular momentum of the binary system as a function of the
coordinate distance $d$, along the evolutionary sequence
$M_{\rm B} = 1.85 \, M_\odot$.}
\label{f:J_1.85}
\end{figure}
\section{Conclusion and perspectives}
We have computed evolutionary sequences of quasiequilibrium irrotational
configurations of binary stars in general relativity.
The evolution of the central density of each star have been monitored
as the orbit shrinks.
For a compactification ratio $M/R=0.14$, the central density
remains rather constant (with a slight increase, below $0.1\%$)
before decreasing. For a higher compactification ratio $M/R=0.17$ (i.e. stars
closer to the maximum mass configuration), a very small density increase
(at most $0.3\%$) is observed before the density
decrease.
It can be thus concluded
that no substantial compression of the stars is found, which means
that no tendency to individually collapse to black hole prior to merger
is observed.
Moreover, the observed density increase remains
within the expected error ($\sim 2\%$, cf. Sect.~\ref{s:assumpt}) induced
by the conformally flat approximation for the 3-metric, so that it cannot
be asserted that this effect would remain in a complete calculation.
No turning point has been found in the binding energy or angular
momentum along evolutionary sequences, which may indicate that these
systems do not have any innermost stable circular orbit (ISCO).
All these results have been obtained for a polytropic EOS with the
adiabatic index $\gamma=2$. We plan to extend them to other values of
$\gamma$ in the near future. We also plan to abandon the conformally
flat approximation for the 3-metric and use the full Einstein equations,
keeping the helicoidal symmetry in a first stage.
\section*{Acknowledgments}
We would like to thank Jean-Pierre Lasota for his constant support and
Brandon Carter for illuminating discussions.
The numerical calculations have been performed on computers
purchased thanks to a special grant from the SPM and SDU departments of CNRS.
\section*{References}
|
\section{INTRODUCTION}
The concept of energy landscapes has played a significant role in
elucidating kinetics of protein folding \cite{Bryngelson,Wolynes}.
An energy landscape can be visualized by using the so called
disconnectivity graphs \cite{Becker} that show patterns of pathways
between the local energy minima of a system. A pathway consists of
consecutive moves that are allowed kinetically. The pathways indicated
in a disconnectivity graph are selected to be those which provide a
linkage at the lowest energy cost among all possible trajectories
between two destinations. Thus, at each predetermined value of a
threshold energy, the local energy minima are represented as divided
into disconnected sets of minima which are mutually accessible through
energy barriers. The local minima which share the lowest energy
barrier are joined at a common node and are said to be a part of a
basin corresponding to the threshold.
The disconnectivity graphs have proved to be useful tools to elucidate
the energy landscape of a model of a short peptide \cite{Becker} and
of several simple molecular systems. In particular, Wales, Miller, and
Walsh \cite{Wales} have constructed disconnectivity graphs for the
archetypal energy landscapes of a cluster of 38 Lennard-Jones atoms,
the molecule of C$_{60}$, and 20 molecules of water. The work on the
Lennard-Jones systems has been recently extended by Doye {\it et al.}
\cite{Doye}.
The graph for a well folding protein is expected to have an appearance
of a ``palm tree.'' This pattern has a well developed basin of the
ground state and it also displays several branches to substantially
higher lying local energy minima. Such a structure seems naturally
associated with the existence of a folding funnel. The atomic level
studies of the 4-monomer peptide considered by Becker and Karplus
\cite{Becker} yield a disconnectivity graph which suggests that this
expected behavior may be correct. Bad folders are expected to have
disconnectivity graphs similar to either a ``weeping willow'' or a
``banyan tree'' \cite{Becker,Wales} in which there are many competing
low lying energy minima.
We accomplish several tasks in this paper. The first of these, as
addressed in Sec. II, is to construct disconnectivity graphs for two
lattice heteropolymers the dynamics of which have been already studied
exactly \cite{Malte}. One of them is a model of a protein, in the
sense that it has excellent folding properties,
and we shall refer to it as a good folder. The other has very poor
folding properties, i.e. it is a bad
folder and is thus a model of a random sequence of aminoacids. We show
that, indeed, only the good folder has a protein-like disconnectivity
graph.
In Sec. III we study the archetypal energy landscapes corresponding
to small two dimensional (2D) Ising spin systems with the
ferromagnetic and spin glassy exchange couplings. We demonstrate that
disordered ferromagnets have protein-like disconnectivity graphs
whereas spin glasses behave like bad folders. This is consistent with
the concept of minimal structural frustration \cite{frustr}, or
maximal compatibility, that has been introduced to explain why natural
proteins have properties which differ from those characterizing random
sequences of aminoacids. It is thus expected that spin systems which
have the minimal frustration in the exchange energy, i.e. the
disordered ferromagnets, would be the analogs of proteins. In fact,
we demonstrate that the kinetics of ``folding,'' i.e. the kinetics of
getting to the fully aligned ground state of the ferromagnet by
evolving from a random state, depends on temperature, $T$, the way a
protein does. Finding a ground state of a similarly sized spin glass
takes place significantly longer.
The disconnectivity graphs characterize the phase space of a system
and, therefore, they relate primarily to the equilibrium properties --
the dynamics is involved only through a definition of what kinds of
moves are allowed, but their probabilities of being implemented are of
no consequence. Note that even if the disconnectivity graph indicates
a funnel-like structure, the system may not get there if the
temperature is not right. Thus a demonstration of the existence of a
funnel must involve an actual dynamics. In fact, another kind of
connectivity graphs between local energy minima has been introduced
recently precisely to describe the $T$-dependent dynamical linkages
\cite{coarse} in the context of proteins. We shall use the phrase a
``dynamical connectivity graph'' to distinguish this concept from that
of a ``disconnectivity graph'' of Becker and Karplus. The idea behind
the dynamical connectivity graphs is rooted in a coarse grained
description of the dynamics through mapping of the system's
trajectories to underlying effective states. In ref. \cite{coarse},
the effective states are the local energy minima arising as a result
of the steepest descent mapping. In ref. \cite{cells}, the steepest
descent procedure is followed by an additional mapping to a closest
maximally compact conformation. The steepest descent mapping has been
already used to describe glasses \cite{Stillinger} and spin glasses
\cite{Jaeckle} in terms of their inherent, or hidden, valley
structures.
In the dynamical connectivity graphs, the linkages are not uniform in
strength. Their strengths are defined by the frequency with which the
two effective states are visited sequentially during the temporal
evolution. The strengths are thus equal to the transition rates and
they vary significantly from linkage to linkage and as a function of
$T$. An additional characteristic used in such graphs is the fraction
of time spent in a given effective state, without making a transition.
This can be represented by varying sizes of symbols associated with
the state.
In the context of these developments, it seems natural to combine the
two kinds of coarse-graining graphs, equilibrium and dynamical, into
single entities -- the supergraphs. Such supergraphs can be
constructed by placing the information about the $T$-dependent
dynamical linkages on the energy landscape represented by the
disconnectivity graph. This procedure is illustrated in Sec. IV for
the case of the two heteropolymers discussed in Sec. II. The
procedure is then applied to selected spin systems. In each case,
knots of significant dynamical connectivities within the ground state
basin develop around a temperature at which the specific heat has a
maximum. These knots disintegrate on lowering the $T$ if the system
is a spin glass or a bad folder. For good folders and non-uniform
ferromagnets the dynamical linkages within the ground state basin
remain robust.
We hope that this kind of combined characterization, by the
supergraphs, of both the dynamics and equilibrium pathways existing in
many body systems might prove revealing also in the case of other
systems, e.g., such as the molecular systems considered in ref.
\cite{Wales}.
\section{ENERGY LANDSCAPES IN 2D LATTICE PROTEINS}
Lattice models of heteropolymers allow for an exact determination of the
native state, i.e. of the ground state of the system, and are endowed with
a simplified dynamics. These two features have allowed for significant
advancement in understanding of protein folding \cite{Dill}.
Here, we consider two 12-monomer sequences of model heteropolymers, $A$ and
$B$, on a two-dimensional square lattice. These sequences have been
defined in terms of Gaussian contact energies (the mean equal to $-1$
and the dispersion to 1, roughly) in ref. \cite{Malte}. They have been
studied \cite{Malte,coarse} in great details by the master equation
and Monte Carlo approaches. Sequences $A$ and $B$ have been established
to be the good and bad folders respectively. Among the $15\;037$ different
conformations that a 12-monomer sequence can take, 495 are the local
energy minima for sequence $A$ and 496 for sequence $B$. The minima are
either $V$- or $U$-shaped. The $U$-shaped minima are those in which a move
that does not change the energy is allowed, provided there are no
moves that lower the energy. Both kinds of minima arise as a result
of the steepest descent mapping from states generated along a Monte
Carlo trajectory and both kinds are included in the disconnectivity
graphs.
Constructing a disconnectivity graph requires determination of the
energy barriers between each pair of the local energy minima. We do
this through an exact enumeration. We divide the energy scale into
discrete partitions of resolution $\Delta E$ (we consider $\Delta
E$=0.5) and ask between what minima there is a pathway which does not
exceed the threshold energy set at the top of the partition. These
minima can then be grouped into clusters which are disconnected from
each other. Local minima belonging to one cluster are connected by
pathways in which the corresponding barriers do not exceed a threshold
value of energy whereas the local minima that belong to different
clusters are separated by energy barriers which are higher than the
threshold level. At a sufficiently high value of the energy threshold
all minima belong to one cluster. Enumeration of the pathways
involves storing a table of size $15\;037 \times 14$ because each
conformation may have up to 14 possible moves within the dynamics
considered in ref. \cite{Malte}. (16-monomer heteropolymers can also be
studied in this exact way -- within any resolution $\Delta E$.)
Figure 1 shows the resulting disconnectivity graphs for sequence $A$.
For clarity, we show only this portion of the graph which involves the
local minima with energies which are smaller than $-5$ (there are 206
such minima). Throughout this paper, the symbol $E$ denotes energy
measured in terms of the coupling constants in the Hamiltonian and is
thus a dimensionless quantity.
The native state, denoted as NAT in the Figure 1,
belongs to the most dominant valley. One can see that the graph
contains a remarkable ``palm tree'' branch that provides a linkage to
the native state. This branch is a place within which a dynamically
defined folding funnel is expected to be confined to. The large size
of this branch associated with a big energy gap between the native
state and other minima indicates large thermodynamic stability. At low
temperatures, the glassy effects set in and contributions due to
non-native valleys become significant. The local minimum denoted by
TRAP in Figure 1 has been identified in ref. \cite{Malte} as giving
rise to the longest lasting relaxation processes in the limit of $T$
tending to 0.
The disconnectivity tree for sequence $B$ is shown in Figure 2. Again,
only the minima with energies smaller than $-5$ are displayed (there are
203 such minima). In this case, there are several local energy minima
which are bound to compete with the native state. The corresponding
branches have comparable lengths and morphologies. The dynamics is
thus expected not to be confined merely to the native basin. Instead,
the system is bound to be frustrated in terms of what branch to choose
to evolve in. At low $T$'s the valley containing the TRAP
conformation is responsible for the longest relaxation and poor
folding properties.
Other examples of disconnectivity trees for protein related systems have
been recently constructed with the use of Go-like models \cite{Li,Miller}
(in which the aminoacid-aminoacid interactions are restricted to
the native contacts)
and they confirm the general pattern of differences in morphology between
good and bad foldability as illustrated by Figures 1 and 2.
It should be noted that there are many ways to map out the multidimensional
energy landscape of proteins. In particular, extensive energy landscape
explorations for the HP lattice heteropolymers have been done with the use
of the pathway maps \cite{MillerDill,Chan1,Chan2}. The pathway maps show
the actual microscopic paths through conformations. The paths are
enumerated either exactly or statistically, and thus provide a detailed but
implicit representation of the energy landscape. The resulting ``flow
diagrams'' indicate patterns of allowed kinetic connections between actual
conformations, together with the energy barriers involved. They can also be
additionally characterized by Monte Carlo determined probabilities to find
a given path at a temperature under study. In this way, preferable
pathways and important transition states can be identified. This approach
is similar in spirit to the one undertaken by Leopold {\it et al.}
\cite{Leopold}
in which the folding funnel is identified through determination of weights
associated with paths that lead to the native state.
The coarse grained representation of energy landscapes in protein-like
systems through the disconnectivity trees is quite distinct from that
obtained through the pathway maps. The disconnectivity graphs indicate
only the one best path for each pair of the local energy minima by showing
the terminal points and the value of the energy barrier necessary to travel
this path. This reduced information is precisely what allows one to
provide an explicit and essentially automatic visualization of the energy
landscapes.
\begin{figure}
\epsfxsize=3.2in
\centerline{\epsffile{fig01.eps}}
\caption{The disconnectivity graph for the 12-monomer
sequence $A$. The dotted area is shown again in Figure 10 together with
the dynamical connectivities. $N_c$ is a symbolic notation for a
label of an energy minimum, based on computer generated listing.}
\end{figure}
\begin{figure}
\epsfxsize=3.2in
\centerline{\epsffile{fig02.eps}}
\caption{Similar to Figure 1 but for sequence $B$. The
dotted area is shown expanded in Figure 11. }
\end{figure}
The $T$-dependent frequencies of passages between conformations in the
pathway maps give an account of the dynamics in the system. This
information on the dynamics, however, does not easily fit the description
provided by the disconnectivity graphs. The steepest descent mapping to the
local energy minima that we propose here is, on the other hand, a perfect
match.
\section{ENERGY LANDSCAPES IN 2D SPIN SYSTEMS}
We now consider the spin systems. The Hamiltonian is given by $H\;=\;
\sum_{<ij>}J_{ij}S_i S_j$ where $S_i$ is $\pm1$, and the exchange
couplings, $J_{ij}$, connect nearest neighbors on the square lattice.
The periodic boundary conditions are adopted. When studying spin
systems, a frequent question to ask about the dynamics is what are the
relaxation times -- characteristic times needed to establish
equilibrium. Here, however, we are interested in quantities which are
analogous to those asked in studies of protein folding. Specifically,
what is the first passage time $t_0$? The first passage time is defined
as the time needed to come across the
ground state during a Monte Carlo evolution that starts from a random
spin configuration. A mean value of $t_0$ in a set of trajectories
(here, we consider 1000 trajectories for each $T$) will be denoted by
$\left<t_0\right>$ and the median value by $t_g$. $t_g$ is an analogue of the
folding time, $t_f$ of ref. \cite{Malte}. At low temperatures, the
physics of relaxation and the physics of folding essentially agree
\cite{Malte}. At high temperatures, however, the relaxation is fast
but finding a ground state is slow due to a large entropy. Both for
heteropolymers and spin systems the $T$-dependence of the
characteristic first passage time is expected to be $U$-shaped. The
fastest search for the ground state takes place at a temperature
$T_{min}$ at which the $T$-dependence has its minimum.
The $U$-shape dependence of $t_f$ originates in the idea of a low $T$
glassy phase in heteropolymers advocated by Bryngelson and
Wolynes \cite{frustr} within the context of the random energy model.
It was subsequently confirmed in numerical simulations of
lattice models \cite{MillerDill,Socci,Chan2}. This shape is, actually
expected for most disordered systems, including those involving spins.
However, experimentalists measuring spin systems typically would
not ask about the first passage time (at high $T$).
This overall behavior is illustrated in Figure 3 for two $5 \times 5$
spin systems. The Gaussian couplings of zero mean and unit dispersion are
selected for the spin glassy (SG) 2$D$ system. The disordered
ferromagnetic system (DFM) is endowed with the exchange couplings
which are the absolute values of the couplings considered for SG.
Figure 3 shows that $t_g$ does depend on $T$ in the $U$-shaped fashion.
$T_{min}$ for SG and DFM are comparable in values but the ``folding''
times for DFM are more than 4 times shorter than for SG. The times are
defined in terms of the number of Monte Carlo steps per spin.
\begin{figure}
\epsfxsize=3.2in
\centerline{\epsffile{fig03.eps}}
\caption{The main figure shows the $T$-dependence of
$t_g$ -- the median time to find the ground state -- for $5 \times 5$
DFM and SG systems. The top inset compares $t_g$ to $t_0$ on the
logarithmic time scale. The divergence of the two times at low $T$'s
indicates a substantial spreading out of the distribution of $t_0$.
This distribution, $P(t_0)$, is shown in the lower inset for
temperatures corresponding to $T_{min}$. }
\end{figure}
\begin{figure}
\epsfxsize=3.2in
\centerline{\epsffile{fig04.eps}}
\caption{Distribution of energy barriers (highest elevation points) across
trajectories. The solid line is for the $4 \times 4$ DFM system.
It shows barriers for all trajectories which connect local minimum
$\alpha$ to another local minimum $\beta$. The lowest of them,
$E_{\alpha \beta}$, is used as threshold in the disconnectivity graph.
The dotted line is for a $5 \times 5$ DFM and for trajectories which
go from a local energy minimum $\phi$ to the ground state.
The energy barrier $E_{\phi 0}$ is obtained through the approximate
enumeration as described in the text. The other values are obtained
by generating 50000 random connecting trajectories. }
\end{figure}
Figure 3 establishes some of the analogies between the heteropolymers
and the spin systems. We now consider the disconnectivity graphs for
selected $L \times L$ spin systems with $L$=4 and 5. For both system
sizes, the list of the local energy minima is obtained through an
exact enumeration. Determination of an energy barrier between two
minima requires adopting some approximations. Suppose that the two
minima differ by $n$ spins. There are then $n!$ possible trajectories
which connect the two minima, assuming that a) no spin is flipped more
than once, b) no other spins (or ``external'' spins) are involved in a
pathway. These trajectories can be enumerated for $L$=4 but not for
$L$=5. In the latter case we adopt the following additional
approximation. We first identify the $n(n-1)(n-2)(n-3)$ list of
the first four possible steps in any trajectory together with the
highest energy elevation reached during these four steps. We choose
$m$=1500 trajectories which accomplish the smallest elevation. We then
consider the next two-step continuations of the selected trajectories
and among the $m(n-4)(n-5)$ continuations again select $m$ which
result in the lowest elevation, and so on until all $n$ spins are
inverted. The lowest elevation among the final set of the $m$
trajectories is an estimate of the energy threshold used in the
disconnectivity diagram. This approximate method, when applied to the
$L$=4 systems, generates results which agree with the exact
enumeration. Figure 4 shows that our method clearly beats
determination of barriers based on totally random trajectories (but
still restricted to overturning of the $n$ differing spins).
Flipping of the ``external'' spins was found to give rise to an
occasional reduction in the barrier height. We could not, however,
come up with a systematic inclusion of such phenomena in the
calculations and the resulting disconnectivity graphs have barriers
which are meant to be estimates from above. The topology of the graph
is expected to depend little on details of such approximations.
\begin{figure}
\epsfxsize=3.2in
\centerline{\epsffile{fig05.eps}}
\caption{Examples of pathways between three local energy
minima, $\alpha, \beta$, and $\gamma$, in a $4 \times 4$ DFM. The
corresponding spin configurations are shown by arrows. The resulting
disconnectivity graph is shown on the left. }
\end{figure}
In some cases, the barrier for a direct travel from one minimum to
another was found to be higher than when making a similar passage via
an intermediate local energy minimum. An example of this situation is
shown in Figure 5. However, this lack of transitivity, resulting from
the approximate nature of the calculations, does not affect the
disconnectivity graph because the states $\gamma$ and $\beta$ of
Figure 5 are mutually accessible at energy $E_{\beta\gamma}$. Then, at
a higher energy $E_{\alpha\gamma}$, state $\alpha$ is thus also
accessible. If, at this energy level, the system can transfer between
the states $\alpha$ and $\gamma$ then it can also transfer to state
$\beta$.
\begin{figure}
\epsfxsize=3.2in
\centerline{\epsffile{fig06.eps}}
\caption{The disconnectivity graphs for the FM and DFM'
$4\times 4$ systems, as defined in the text. The arrows in the boxes
show examples of the corresponding spin configurations. The numbers
indicate the degree of degeneracy and the numbers in brackets indicate
numbers of distinct geometries for the inverted domains to take at the
energy considered. For a uniform antiferromagnetic system the
disconnectivity graph looks qualitatively similar to the one
characterizing FM but the ground state is the only $V$-shaped local
energy minimum of the system.}
\end{figure}
We now present specific examples of disconnectivity graphs for several
distinct spin systems.
Figure 6 shows the case of a $4\times 4$ uniform ferromagnet (FM).
The energy landscape of the FM is not analogous to that of a protein
because uniform exchange couplings generate states with high degrees
of degeneracies. These degeneracies can be split either by a
randomization. Figure 6 also shows a graph for a $L=4$ DFM' system in
which the $J_{ij}$'s are random numbers from the [0.9,1.1] interval --
this is the case of a small perturbation away from the uniform FM.
The graph for DFM' has an overall appearance like the one for FM
except for the lack of a high energy linkage to a set of state which
cease to be minima. Another difference is the disappearance of all
remaining $U$-shaped minima and formation of new true minima at somewhat
spread out energies. In the uniform $L=4$ ferromagnet, there are five
$V$-shaped energy minima: one is the ground state and the other four
higher energy states are degenerate. In addition, there are 346
states which are the $U$-shaped energy minima. An example of what
happens in a $U$-shaped minimum is shown in Figure 7. Here, the system
can move between the 3- and 4-spin domains without a change in the
energy. The 4-spin domain forms a $U$-shaped minimum but the 3-spin
state is not a minimum because there is a move to a lower energy
state. Only the 4-spin domain states would be shown in the
disconnectivity graph.
\begin{figure}
\epsfxsize=3.2in
\centerline{\epsffile{fig07.eps}}
\caption{Examples of spin configurations in the $4\times 4$ FM system.}
\end{figure}
\begin{figure}
\epsfxsize=3.2in
\centerline{\epsffile{fig08.eps}}
\caption{The disconnectivity graphs for the $4 \times4$
spin glassy systems: with the Gaussian (SG') and with the $\pm 1$
couplings (SG $\pm$). The numbers correspond to the number of
$U$-shaped minima at an energy shown in the graph.}
\end{figure}
Figure 8 shows the disconnectivity graphs for two $L=4$ spin glassy
systems. The right hand panel shows the case of $J_{ij}= \pm 1$. The
left hand panel shows a spin glass (SG') with the exchange couplings
which are randomly positive or negative and with their magnitudes
coming from the interval [0.9,1.1] -- this is the random sign
counterpart of the DFM' system. In both spin glassy systems of Figure
8 the allocation of signs to the couplings is identical. In the $\pm
1$ case, all minima, including the degenerate ground state, are
$U$-shaped. The SG' system, on the other hand, has a graph with an
overall structure akin to that corresponding to the $\pm 1$ system
with one important difference: the ground state is not degenerate and
thus the ground state basin splits into several competing valleys.
The differences between the good and bad spin ``folders'' amplify as the
system size is increased. As an illustration, Figure 9 shows the
disconnectivity graphs for the $5 \times 5$ DFM and SG -- with the
Gaussian couplings. The DFM systems has a very stable and well
developed valley corresponding to the ground state whereas the SG
system has many competing valleys. Thus indeed, DFM is a spin analogue
of a protein whereas SG is an analogue of a random sequence of
aminoacids.
\begin{figure}
\epsfxsize=3.2in
\centerline{\epsffile{fig09.eps}}
\caption{The disconnectivity graphs for the $5 \times 5$
DFM and SG systems (the top panels) and the corresponding
representation of the energy landscapes (the bottom panels).}
\end{figure}
The disconnectivity graphs can be represented in a form that gives a
better illusion of an actual landscape, as shown in the bottom panels
of Figure 9. The lines shown there connect the local energy minima to
their energy barriers and then to the next minimum, and so on, forming
an envelope of the original graph. This form is less cluttered and
will be used in Sec. V. This envelope representation shows merely
the smallest scale variations in energy and omits passages with
large barriers.
\section{DYNAMICAL CONNECTIVITY GRAPHS FOR LATTICE HETEROPOLYMERS}
We now construct the supergraphs for the lattice heteropolymers
discussed in Sec. II. The strengths of the dynamical linkages have been
already determined in ref. \cite{coarse} at several temperatures.
Here, however, we plot the linkages on the graphs that represent the
energy landscapes, i.e. we rearrange the labels associated with the
local energy minima. We discuss only the case of $T=T_{min}$ which is
equal to 1.0 for both sequences $A$ and $B$.
\begin{figure}
\epsfxsize=3.2in
\centerline{\epsffile{fig10.eps}}
\caption{Dynamical connectivity graph for sequence $A$ at
$T$=1.0 plotted against the background of the disconnectivity graph.
The dynamical linkages are restricted to the dotted region of Figure 1
and only this portion of the disconnectivity graph is shown.}
\end{figure}
\begin{figure}
\epsfxsize=3.2in
\centerline{\epsffile{fig11.eps}}
\caption{Similar to Figure 10 but for sequence $B$.}
\end{figure}
Figures 10 and 11 shows the supergraphs for sequences $A$ and $B$
respectively. The sizes of the circles are proportional to an
occupancy of the minimum during the folding time. Similarly, the
thicknesses of the lines connecting the circles are proportional to
the connectivity (the linking frequency) between them. For clarity,
we do not show connectivities which account for less than $1\%$ of all
combined dynamical connectivities. The disconnectivity graphs
themselves are drawn in dotted lines. All relevant dynamics is
confined to these portion of the of the disconnectivity graphs which
were marked, in Figures 1 and 2, by the dotted lines and are now
magnified in Figures 10 and 11.
An inspection of the supergraphs clearly shows differences between the
two sequences. Sequence $A$ has many inter-valley linkages but the
linkages to the native basin, and the occupancies of conformations
within that basin, are substantial. These are manifestations of a
fast folding dynamics. For sequence $B$, on the other hand, the linkages
tend to wither uncooperatively in multiple valleys. In addition, the
combined occupancies away from the native valley outweigh the
dynamical effects within the valley. On lowering the temperature,
linkages in various valleys become disconnected and tend to avoid the
native valley more and more, as discussed in ref. \cite{coarse}.
\section{DYNAMICAL CONNECTIVITY GRAPHS FOR SPIN SYSTEMS}
We now generate dynamical linkages for two spin systems, $L=5$ DFM and
SG of Sec. III, and place them on the plots of the energy landscape.
The ``envelope'' form of the representation of the landscape is chosen
here, mostly for esthetic reasons. The connectivities are determined
based on 200 Monte Carlo trajectories of a fixed length of 5000 steps
per spin. The duration of these trajectories exceeds the ``folding
time'' many times, at the temperatures studied, and thus the
connectivities displayed refer to the essentially equilibrium
situations (the equilibrium dynamics for heteropolymers $A$ and $B$ is
illustrated in ref. \cite{coarse}). The connectivity rates were
updated any time (in terms of single spin events and not in terms of
steps per spin) there is a transition from a local energy minimum to a
local energy minimum, after the steepest descent mapping. Again, the
$1\%$ display cutoff has been implemented when making the figure.
The main parts of Figures 12 and 13 show the supergraphs obtained at a
temperature which corresponds to the $T$-location of the peak in
specific heat. These temperatures, 1.8 for DFM and 1.4 for SG, are
also close to $T_{min}$. The insets show the dynamically relevant
parts of the energy landscape at lower temperatures. For the DFM, the
dynamics becomes increasingly confined to the ground state basin when
the temperature is reduced. On the other hand, for the SG, the
dynamics in the ground state basin becomes less and less relevant,
with a higher local energy minimum absorbing the majority of moves.
This is indeed what happens with bad folding heteropolymers.
If we restrict counting of the transition rates only to the ``folding
stage,'' i.e. till the ground state is encountered, the qualitative
look of the supergraph for $T$ close to $T_{min}$ is as in the
equilibrium case. The states involved are mostly the same but there
is, by definition, only one link to the ground state per trajectory.
The dynamical connectivity graphs in 3D $10\times 10\times 10$ DFM
systems are qualitatively similar to the 2D graphs but the
underlying disconnectivity graphs are harder to display due to a
substantially larger number of the energy minima.
\begin{figure}
\epsfxsize=3.2in
\centerline{\epsffile{fig12.eps}}
\caption{The dynamical-equilibrium supergraph for the
$5\times 5$ DFM system at $T$=1.8. The inset shows the portion
that is relevant at $T$=0.9.}
\end{figure}
\begin{figure}
\epsfxsize=3.2in
\centerline{\epsffile{fig13.eps}}
\caption{The supergraph for the $5\times 5$ SG system at
$T$=1.4. The insets show the portions which are relevant at two lower
temperatures. The ground state configuration is shown at the top.}
\end{figure}
In this paper we have pointed out the existence of many analogies
between protein folding and dynamics of spin systems. These analogies
have restrictions. For instance the simple Ising spin systems in 3$D$
have continuous phase transitions, in the thermodynamic limit, and not
the first-order-like that are expected to characterize large proteins
\cite{Gutin}. This difference, however, is not crucial in the case of
small systems. More accurate spin analogs of proteins, with the first
order transition, can be constructed but the object of this paper was
to discuss the basic types of spin systems.
On the other hand, it should be pointed out that these analogies are
also more extensive. Consider, for instance, the Thirumalai
\cite{Thirumalai} criterion for good foldability of proteins. The
criterion considers two quantities: the specific heat and the
structural susceptibility of a heteropolymer. The latter is a measure
of fluctuations in the structural deviations away from the native
state. Both quantities have peaks at certain temperatures. The
criterion specifies that if the two temperatures coincide a
heteropolymer is a good folder. This is quite similar to what happens
in uniform and disordered 3D ferromagnets: the peaks (singularities)
in magnetic susceptibility and specific heat are located at the same
critical temperature. On the other hand, in spin glasses, the broad
maximum in the specific heat is located at a temperature which is
substantially above the freezing temperature associated with the cusp
in the susceptibility. Also in this sense then, spin glasses behave
like bad folders.
The coarse-graining supergraphs that analyse dynamics in the context
of the system's energy landscape may become a valuable tool to
understand complex behavior of many body systems.
\section*{ACKNOWLEDGMENTS}
This work was supported by KBN (Grants No. 2P03B-025-13 and
2P03B-125-16). Fruitful discussions with Jayanth R. Banavar are
appreciated.
\vspace{0.5cm}
|
\section{Introduction}
The search for supersymmetry (SUSY) is one of the main
goals of the present and future colliders. In particular an $e^{+}
e^{-}$ linear collider will be an excellent discovery machine for SUSY
particles \cite{JLC}. Among them the charginos --
mixtures of W-inos and charged higgsinos -- are of particular
interest. They belong to the lightest SUSY particles and
are produced with large cross sections if kinematically accessible. Their
properties depend on the SU(2) gaugino mass $M$, the higgsino mass
parameter $\mu$ and the ratio $\tan\beta=v_2/v_1$ of the
vacuum expectation values of the two neutral Higgs fields.
A number of studies addressed the determination of these fundamental
SUSY parameters from chargino production at $e^{+} e^{-}$ colliders
\cite{feng,etal,tsukamoto}.
Recently a procedure was proposed to
determine the parameters $M$, $\mu$ and $\tan\beta$
independently from the decay dynamics \cite{choi}.
For this, however, one has to assume that the selectron sneutrino
mass $m_{\tilde{\nu}_e}$ is already known, e.g. from sneutrino pair
production.
Since for gaugino-like charginos
the cross section sensitively depends on the sneutrino
mass the possibility to measure $m_{\tilde{\nu}_e}$ in chargino pair
production at an $e^{+}e^{-}$ collider has also been discussed
\cite{feng,tsukamoto}.
In this paper we study the prospects for constraining
$m_{\tilde{\nu}_e}$ with suitably polarized beams
by combining information from the cross
section
\begin{equation}
\sigma_{e^{-}}=
\sigma(e^{+}e^{-}\to\tilde{\chi}^{+}_1 \tilde{\chi}^{-}_1)\times
BR(\tilde{\chi}^{-}_1\to\tilde{\chi}^0_1 e^{-}
\bar{\nu}_e)\label{sigma}
\end{equation}
and the forward--backward asymmetry (FB--asymmetry)
\begin{equation}
A_{FB}=\frac{\sigma_{e^{-}}(\cos\Theta_{-}>0)-\sigma_{e^{-}}(\cos\Theta_{-}<0)}
{\sigma_{e^{-}}(\cos\Theta_{-}>0)+\sigma_{e^{-}}(\cos\Theta_{-}<0)}\label{afb}
\end{equation}
of the decay lepton $e^{-}$ from the decay
$\tilde{\chi}^{-}_1\to\tilde{\chi}^0_1 e^{-} \bar{\nu}_e$. In eq.(\ref{afb})
$\Theta_{-}$ is the angle between the incoming electron beam and the outgoing
$e^{-}$ in the laboratory system.
For constraining $m_{\tilde{\nu}}$ the simultaneous polarization of both
beams turns out to be very useful \cite{cracow98}.
In Section~2 the general formalism is presented and in Section~3
the numerical results are discussed.
\section{General formalism}
\vspace{-.3cm}
The production process
$e^{+} e^{-} \to \tilde{\chi}^{+}_1 \tilde{\chi}^{-}_1$
contains contributions from $\gamma$, $Z^0$ and $\tilde{\nu}_e$ exchange
and the decay process, $\tilde{\chi}^{-}_1 \to
\tilde{\chi}^0_1 \ell^{-} \bar{\nu}_e$
($\tilde{\chi}^{+}_1 \to
\tilde{\chi}^0_1 \ell^{+} \nu_e$)
contains contributions from $W$, $\tilde{\ell}_L$ and
$\tilde{\nu}_e$ exchange.
These processes have been studied in \cite{moor} properly taking into account
the spin correlations between chargino production and decay.
In the following the amplitude for the production (decay)
of $\tilde{\chi}^{-}_1$ with
helicity $\lambda_1$ is denoted by $P^{\lambda_1}$ ($D_{\lambda_1}$).
All other helicity indices are suppressed.
Then the amplitude of the combined process is
\begin{equation}
T=\Delta_{\tilde{\chi}^{-}_{1}} \sum_{\lambda_1}
P^{\lambda_1} D_{\lambda_1}.\label{eq_1}
\end{equation}
In eq.~(\ref{eq_1})
$\Delta_{\tilde{\chi}^{-}_{1}}=1/[p^2_{\tilde{\chi}^{-}_{1}}
-m^2_{\tilde{\chi}^{-}_{1}}+i m_{\tilde{\chi}^{-}_{1}}
\Gamma_{\tilde{\chi}^{-}_{1}}]^{-1}$, $p^2_{\tilde{\chi}^{-}_1}$,
$m_{\tilde{\chi}^{-}_{1}}$ and
$\Gamma_{\tilde{\chi}^{-}_{1}}$
denote the propagator, the
four--momentum squared, the mass and the width of
$\tilde{\chi}^{-}_1$. For this propagator we use the narrow width
approximation.
The amplitude squared
$|T|^2=|\Delta_{\tilde{\chi}^{-}_{1}}|^2 \sum_{\lambda_1 \lambda^{'}_1}
\rho^{\lambda_1 \lambda_1'}_P
\rho^D_{\lambda_1' \lambda_1} \label{N}$
is thus composed of the
unnormalized spin density matrix
$\rho^{\lambda_1 \lambda_1'}_P=P^{\lambda_1} P^{\lambda_1' *}$
of $\tilde{\chi}^{-}_1$ and
the decay matrix
$\rho^{D}_{\lambda_1' \lambda_1}=D_{\lambda_1} D_{\lambda_1'}^{*}$.
Interference terms between
various helicity amplitudes preclude factorization into a production
factor $\sum_{\lambda_1} |P^{\lambda_1}|^2$ times a decay factor
$\overline{\sum}_{\lambda_1} |D_{\lambda_1}|^2$.
For the general case $e^{+}e^{-}\to \tilde{\chi}^{+}_i
\tilde{\chi}^{-}_j$ $(i, j=1, 2)$,
$\tilde{\chi}^{+}_i\to\tilde{\chi}^0_k \ell^{+} \nu_{\ell}$,
$\tilde{\chi}^{-}_j\to\tilde{\chi}^0_l \ell^{-} \bar{\nu}_{\ell}$ $(k,
l=1,\ldots,4)$
the complete analytical
formulae for polarized beams with full spin correlations between
production and decay are given in \cite{moor}.
\section{Numerical Results}
Our study is embedded in the Minimal Supersymmetric Standard
Model (MSSM) for two representative
scenarios with practically the same mass of the
lighter chargino and the lightest neutralino $\tilde{\chi}^0_1$,
respectively. The two scenarios differ,
however, significantly in the mixing character of the chargino.
The chargino mass eigenstates $\tilde{\chi}_i={\chi^+_i \choose
\chi^{-}_i}$ are defined by $\chi^{+}_i=V_{i1} w^{+}+V_{i2} h^{+}$
and $\chi^{-}_i=U_{i1} w^{-}+U_{i2} h^{-}$. Here $w^{\pm}$ and
$h^{\pm}$ are the two-component spinor fields of the W-ino and the charged
higgsinos, respectively.
Furthermore $U_{ij}$ and $V_{ij}$ are the elements of the $2\times 2$
matrices which diagonalize the chargino mass matrix.
Below we give the W-ino and higgsino components of the two-component spinor
field $\chi^{-}_1$. Similarly we give the components of the neutralino
$\tilde{\chi}^0_1$ in the B-ino -- W-ino basis
$(\tilde{B}|\tilde{W}|\tilde{H}^0_a|\tilde{H}^0_b)$. For details see
\cite{haber,bartl86}.
\setlength{\mathindent}{0cm}
\[
\mbox{\bf Scenario~A}:\quad M=152\mbox{~GeV}, M'=78.7\mbox{~GeV},
\mu=316\mbox{~GeV}, \tan\beta=3.
\]
Here we have used the GUT relation $M'=\frac{5}{3} M \tan^2 \Theta_W$.
This set of parameters is inspired by the low $\tan\beta$ mSUGRA
reference scenario specified in \cite{blair}.
The eigenstates and the masses of the lighter chargino and of the LSP
$\tilde{\chi}^0_1$ are
\begin{eqnarray*}
\chi^{-}_1&=&(-0.91|+0.42)\quad\mbox{and}\quad
m_{\tilde{\chi}_1^{-}}=128\mbox{~GeV},\\
\tilde{\chi}^0_1&=&(-0.97|+0.15|-0.15|-0.15)\quad\mbox{and}\quad
m_{\tilde{\chi}^0_1}=71\mbox{~GeV}.\\
\end{eqnarray*}
\[
\mbox{\bf Scenario~B}:\quad M=400\mbox{~GeV}, M'=95\mbox{~GeV},
\mu=145\mbox{~GeV}, \tan\beta=3.
\]
Here the parameters are chosen thus that the chargino is higgsino-like
with only a small gaugino component. To obtain nearly the same masses as in
Scenario~A and an as
large as possible higgsino component for $\tilde{\chi}^0_1$
we take $M$ and $M'$ as independent
parameters and do not use the GUT relation between them.
The eigenstates and masses are
\begin{eqnarray*}
&&\chi^{-}_1=(-0.19|+0.98)\quad\mbox{and}\quad
m_{\tilde{\chi}_1^{-}}=129\mbox{~GeV},\\
&&\tilde{\chi}^0_1=(-0.82|+0.11|-0.45|-0.33)\quad\mbox{and}\quad
m_{\tilde{\chi}^0_1}=71\mbox{~GeV}.\\
\end{eqnarray*}
In both scenarios the LSP $\tilde{\chi}^0_1$ has a dominating B-ino component.
We vary the sneutrino mass $m_{\tilde{\nu}_e}$ between 80~GeV and
1000~GeV.
The mass of the left selectron $m_{\tilde{e}_L}$ is determined
by the relation
\begin{equation}
\phantom{Scenario~A: \quad} m^2_{\tilde{e}_L}-m^2_{\tilde{\nu}_e}
=-M_W^2\cos 2\beta.
\end{equation}
This sum rule is independent of any assumption about the gaugino mass
parameters and on the assumption of a universal scalar mass $m_0$
\cite{martin}.
The influence of spin correlations between production and decay on the
decay angular distributions is strongest near threshold.
Therefore numerical results are presented for energies near threshold
$\sqrt{s}=270$~GeV and for
$\sqrt{s}=500$~GeV.
We have calculated for longitudinally polarized beams
the cross section $\sigma_{e^{-}}$, eq.~(\ref{sigma}), and the FB--asymmetry
$A_{FB}$, eq.~(\ref{afb}), of the decay electron.
For the polarization of the $e^{-}$ beam we assume $P_{-}=\pm85\%$
and for the polarization of the $e^{+}$ beam we take $P_{+}=\pm60\%$.
Simultaneous polarization of both beams is not only useful to enlarge
$\sigma_{e^{-}}$ but it can also have strong influence on $A_{FB}$
according to the mixing character of the charginos.
Since the sneutrino couples only to left-handed electrons the
polarization $P_{-}<0$ ($P_{+}>0$) for the $e^{-}$($e^{+}$) beam is
favourable for studying its properties.
\subsection{Cross sections}
For the gaugino-like scenario~A and $\sqrt{s}=270$~GeV the cross section
$\sigma_{e^{-}}$ is shown in
Fig.~1.
For $m_{\tilde{\nu}_e}$ between 80~GeV and 400~GeV
it is very sensitive on $m_{\tilde{\nu}_e}$. The deep dip at
$m_{\tilde{\nu}_e}\approx 130$~GeV
is due to the destructive interferences between gauge boson and
slepton exchange \cite{bartl86,bartl92}. For
$m_{\tilde{\nu}_e}<m_{\tilde{\chi}^{-}_1}$ the cross section
increases owing to the two-body decay
$\tilde{\chi}^{-}_1\to\tilde{\nu}_e e^{-}$. However, for
$\sqrt{s}=270$~GeV this effect is suppressed by destructive
interference effects.
For higher energy $\sqrt{s}=500$~GeV, Fig.~2, the destructive
$Z^0$--$\tilde{\nu}_e$ interference
in production is shifted to higher $m_{\tilde{\nu}_e}$ and
the two-body decay threshold is more apparent.
For values of $m_{\tilde{\nu}_e}>200$~GeV
the cross section $\sigma_{e^{-}}$ for
$\sqrt{s}=500$~GeV is smaller than for $\sqrt{s}=270$~GeV.
Beam polarization has strong influence on the magnitude of $\sigma_{e^{-}}$.
For $P_{-}=-85\%$ and $P_{+}=+60\%$ it increases
by a factor of about 3, nearly independent of $\sqrt{s}$ and of
$m_{\tilde{\nu}_e}$, Fig.~1 and Fig.~2.
For the opposite polarization configuration, $P_{-}=+85\%, P_{+}=-60\%$,
it would decrease by a factor of 0.06.
If only the electron beam is polarized
$\sigma_{e^{-}}$
increases for $P_{-}=-85\%$ by a factor of 1.85 and for
$P_{-}=+85\%$ it decreases by a factor of 0.15.
In Fig.~3 and Fig.~4
we show $\sigma_{e^{-}}$ for the higgsino-like scenario~B.
Since the dominating higgsino components do not couple to sneutrinos,
the small gaugino admixture causes only a weak dependence on
$m_{\tilde{\nu}_e}$, Fig.~3.
The destructive interference between $Z^0$ and $\tilde{\nu}_e$ exchange
is suppressed by the dominating higgsino component.
For $m_{\tilde{\nu}_e}>m_{\tilde{\chi}^{-}_1}$ the cross section
is practically independent of
$m_{\tilde{\nu}_e}$.
Only for $m_{\tilde{\nu}_e}<m_{\tilde{\chi}^{-}_1}$ it shows a strong
dependence on the sneutrino mass due to
the two-body decay
$\tilde{\chi}^{-}_1\to\tilde{\nu}_e e^{-}$.
For higher energy, $\sqrt{s}=500$~GeV, Fig.~4,
$\sigma_{e^{-}}$ is smaller. The dependence on $m_{\tilde{\nu}_e}$,
however, is essentially unchanged.
Also for higgsino-like charginos
beam polarization has
a strong influence on the magnitude of $\sigma_{e^{-}}$.
For $P_{-}=-85\%$ and $P_{+}=+60\%$ and for both values of $\sqrt{s}$,
Fig.~3 and Fig.~4, $\sigma_{e^{-}}$ increases
by a factor of about 2.5, which is somewhat smaller as for
the gaugino-like scenario~A.
For the opposite polarization, $P_{-}=+85\%, P_{+}=-60\%$, $\sigma_{e^{-}}$
decreases by a factor of 0.5.
If only the $e^{-}$ beam is polarized with $P_{-}=-85\%$($P_{-}=+85\%$)
the cross section increases(decreases)
by a factor of 1.6(0.4).
\subsection{Forward--backward asymmetry of the decay lepton}
In this section we study the FB--asymmetry $A_{FB}$
of the decay electron, eq.~(\ref{afb}),
for $m_{\tilde{\nu}_e}$ between 80~GeV and
1000~GeV. Since the angular distribution
of the chargino decay products is determined by the polarization of
the decaying chargino, the FB--asymmetry of the decay lepton
strongly depends on the
spin correlations between production and decay.
The dependence of $A_{FB}$ on $m_{\tilde{\nu}_e}$, on the beam polarization
and on $\sqrt{s}$ is a result of the complex interplay
between production and decay.
In scenario~A for $\sqrt{s}=270$~GeV the dependence of $A_{FB}$
on $m_{\tilde{\nu}_e}$
is shown in Fig.~5. For high $m_{\tilde{\nu}_e}>400$~GeV
it is about $8\%$ and is nearly independent of $m_{\tilde{\nu}_e}$.
For $m_{\tilde{\nu}_e}<400$~GeV, however, $A_{FB}$ shows
a complex and interesting mass dependence.
The destructive $Z^0$--$\tilde{\nu}_e$ interference
in the production
causes large $A_{FB}\approx 30\%$. Since this destructive
interference effect vanishes at about $m_{\tilde{\nu}_e}\approx 200$~GeV
the FB--asymmetry decreases for higher values of $m_{\tilde{\nu}_e}$.
The deep dip at $m_{\tilde{\nu}_e}\approx 140$~GeV is due to destructive
$W$--$\tilde{\nu}_e$ ($W$--$\tilde{e}$) interference in the decay process
$\tilde{\chi}^{-}_1\to \tilde{\chi}^0_1 e^{-} \bar{\nu}_e$. For
$m_{\tilde{\nu}_e}\le m_{\tilde{\chi}^{-}_1}$ the FB--asymmetry
strongly increases up
to about $35\%$, owing to the two-body decay
$\tilde{\chi}^{-}_1\to\tilde{\nu}_e e^{-}$.
For $\sqrt{s}=500$~GeV, Fig.~6,
large values of $A_{FB}$ up to $50\%$ can be reached in scenario~A.
The $Z^0$--$\tilde{\nu}_e$ interference effect in production
is shifted to higher $m_{\tilde{\nu}_e}$ so that just
for $m_{\tilde{\nu}_e}\ge 250$~GeV $A_{FB}$ decreases
from $48\%$ to $10\%$ for $m_{\tilde{\nu}_e}=1000$~GeV.
For $\sqrt{s}=500$~GeV the $Z^0$--$\tilde{\nu}_e$ interference
in the decay process is less important because
for higher energies the characteristics of the production process
is more and more dominating.
Therefore at $m_{\tilde{\nu}_e}\approx 140$~GeV $A_{FB}$
decreases only to about $43\%$ in contrast to the deep dip at
$m_{\tilde{\nu}_e}\approx 140$~GeV for $\sqrt{s}=270$~GeV.
Again, for $m_{\tilde{\nu}_e}\le
m_{\tilde{\chi}^{-}_1}$
the two-body decay
$\tilde{\chi}^{-}_1\to\tilde{\nu}_e e^{-}$ involves large
$A_{FB}\approx 50\%$.
Note that in the gaugino-like
scenario~A $A_{FB}$ is nearly independent on beam polarization.
This emerges as a consequence of the fact that
in scenario~A the beam polarization
changes the magnitude of $\sigma_{e^{-}}$ by a factor which is
nearly independent
of $m_{\tilde{\nu}_e}$. This factor cancels
in $A_{FB}$, eq.~(\ref{afb}).
Only for $P_{-}>0$ a small polarization effect can be observed, see
Fig.~5 and Fig.~6.
In the higgsino-like scenario~B, Fig.~7, the FB--asymmetry for
$\sqrt{s}=270$~GeV
is much smaller than in scenario~A, $A_{FB}\approx 5\%$
for $m_{\tilde{\nu}_e}\ge m_{\tilde{\chi}^{-}_1}$.
Only for $m_{\tilde{\nu}_e}\le m_{\tilde{\chi}^{-}_1}$
when the two-body decay
$\tilde{\chi}^{-}_1\to e^{-} \tilde{\nu}_e$ is open
large $A_{FB}\approx 33\%$ can be reached for unpolarized
beams.
For $\sqrt{s}=500$~GeV $A_{FB}$ decreases,
$A_{FB}\le 5\%$ for $m_{\tilde{\nu}_e}\ge m_{\tilde{\chi}^{-}_1}$
and $A_{FB}\approx 18\%$ for
$m_{\tilde{\nu}_e}\le m_{\tilde{\chi}^{-}_1}$, Fig.~8.
In this higgsino-like scenario~B the FB--asymmetry sensitively
depends on the beam polarization.
For $P_{-}=-85\%, P_{+}=+60\%$
the FB--asymmetry
increases for $m_{\tilde{\nu}_e}\le m_{\tilde{\chi}^{-}_1}$
up to $48\%$
for $\sqrt{s}=270$~GeV and up to $28\%$ for $\sqrt{s}=500$~GeV.
If only the electron beam is polarized
with $P_{-}=+85\%$,
one even obtains for $m_{\tilde{\nu}_e}\le m_{\tilde{\chi}^{-}_1}$
large negative FB--asymmetries, $A_{FB}\approx -20\%$($A_{FB}\approx -12\%$),
for $\sqrt{s}=270$~GeV($\sqrt{s}=500$~GeV).
\subsection{Constraining $m_{\tilde{\nu}_e}$}
In this section we study the prospects to constrain
$m_{\tilde{\nu}_e}$ by measuring $\sigma_{e^{-}}$ and
$A_{FB}$ with different beam polarizations.
We assume that $\tan\beta$, $m_{\tilde{\chi}^{-}_1}$ and
$m_{\tilde{\chi}^0_1}$ are already known and we choose as an example the
gaugino-like scenario~A and the higgsino-like scenario~B.
One obtains the largest cross sections for polarizations
$P_{-}<0$,~$P_{+}>0$. We therefore assume that $\sigma_{e^{-}}$
has been measured for beam
polarizations $P_{-}=-85\%$, $P_{+}=+60\%$ and for energies
$\sqrt{s}=270$~GeV and $\sqrt{s}=500$~GeV with an error of 5\%.
We study also the correponding $A_{FB}$ for two combinations of
beam polarizations, $P_{-}=-85\%$, $P_{+}=+60\%$ and
$P_{-}=+85\%, P_{+}=0\%$.
For the second case we use unpolarized positrons since also
switching the $e^{+}$ polarization gives extremely small rates.
\begin{center}
{\it 3.3.1. Constraining $m_{\tilde{\nu}_e}$ for $\sqrt{s}=270$~GeV}
\end{center}
We assume that for $\sqrt{s}=270$~GeV and $P_{-}=-85\%$,
$P_{+}=+60\%$ a cross section of
$\sigma_{e^{-}}=180 fb\pm 5\%$
has been measured.
As can be seen from Fig.~1 for scenario~A
and from Fig.~3 for scenario~B this cross section
is compatible with two regions for $m_{\tilde{\nu}_e}$
in either case.
In scenario~A the cross section
is compatible with
small values of $m_{\tilde{\nu}_e}$ between 85~GeV and 89~GeV
and with high $m_{\tilde{\nu}_e}$ between 240~GeV and 255~GeV,
Fig.~1.
For the case of low $m_{\tilde{\nu}_e}$ the FB--asymmetry is between
29\% and 32\% and for high $m_{\tilde{\nu}_e}$
between 19\% and 22\%, Fig.~5.
Accordingly $A_{FB}$ allows to discriminate between the two regions for
$m_{\tilde{\nu}_e}$.
Changing the polarization to $P_{-}=+85\%$, $P_{+}=0\%$, the
FB--asymmetry gives no additional constraints for $m_{\tilde{\nu}_e}$, Fig.~5.
In scenario~B the cross section is
compatible with small $m_{\tilde{\nu}_e}$ between
128~GeV and 140~GeV near the two-body decay threshold and with
high $m_{\tilde{\nu}_e}>250$~GeV, Fig.~3.
For the case of low $m_{\tilde{\nu}_e}$ the FB--asymmetry is between
11\% and 29\% and for high $m_{\tilde{\nu}_e}$
$A_{FB}$ is less than $5\%$, Fig.~7.
Again $A_{FB}$ allows to discriminate between the two regions for
$m_{\tilde{\nu}_e}$.
In our examples with $P_{-}=-85\%$, $P_{+}=+60\%$
the values of the FB--asymmetry for the low mass region of
scenario~B (11\%--29\%) overlaps with that for the low mass
region of scenario~A (29\%--32\%) as well as with that of the high mass
region of scenario~A (19\%--22\%).
However measuring $A_{FB}$ for a suitably
polarized electron beam and an unpolarized positron beam allows to
discriminate between scenario~A and scenario~B.
For $P_{-}=+85\%$, $P_{+}=0\%$ the FB--asymmetry in the low mass region of
scenario~B is negative between $-5$\% and $-16\%$, whereas for scenario~A
the respective regions for $A_{FB}$ are unchanged, compare Fig.~5 and Fig.~7.
\begin{center}
{\it 3.3.2. Constraining $m_{\tilde{\nu}_e}$ for $\sqrt{s}=500$~GeV}
\end{center}
In the same manner we study
$\sigma_{e^{-}}$ and $A_{FB}$ for $\sqrt{s}=500$~GeV.
We assume that
for polarized beams with $P_{-}=-85\%$, $P_{+}=+60\%$ the cross section
$\sigma_{e^{-}}=130 fb\pm 5\%$ has been measured.
As can be seen from Fig.~2 for scenario~A and Fig.~4
for scenario~B this cross section
is again compatible with two regions for $m_{\tilde{\nu}_e}$ in
either case.
In scenario~A $\sigma_{e^{-}}$
is compatible with small
values of $m_{\tilde{\nu}_e}$ between $161$~GeV and $175$~GeV and
with high values of $m_{\tilde{\nu}_e}$ between $420$~GeV and $450$~GeV,
Fig.~2.
For small $m_{\tilde{\nu}_e}$ the
FB--asymmetry is about $45\%$ whereas
for high $m_{\tilde{\nu}_e}$
it is between $32\%$ and $35\%$, clearly
distinguishable from the case of
low $m_{\tilde{\nu}_e}$, Fig.~6. As
compared to the case of lower
energies, $\sqrt{s}=270$~GeV, the values of
$A_{FB}$ are higher and restricted to
smaller regions.
Changing the beam polarization to $P_{-}=+85\%$, $P_{+}=0\%$
the FB--asymmetry gives practically
no additional constraints for $m_{\tilde{\nu}_e}$, Fig.~6.
In scenario~B $\sigma_{e^{-}}$ is compatible with
small $m_{\tilde{\nu}_e}$ in the region between $122$~GeV and $128$~GeV
near the two-body decay threshold and with
high $m_{\tilde{\nu}_e}>750$~GeV, Fig.~4.
For low $m_{\tilde{\nu}_e}$ the FB--asymmetry
is between $10\%$ and $28\%$ whereas for high values of $m_{\tilde{\nu}_e}$
it is rather small, $A_{FB}<3\%$.
These two regions for $m_{\tilde{\nu}_e}$
can clearly be distinguished by
measuring $A_{FB}$, Fig.~8.
Similar as for lower energies, $\sqrt{s}=270$~GeV, the regions for
$A_{FB}$ for the case of high $m_{\tilde{\nu}_e}$ in scenario~A and for the
case of $m_{\tilde{\nu}_e}\approx m_{\tilde{\chi}^{-}_1}$ in scenario~B
are not well separated.
Again a suitably polarized $e^{-}$ beam and an unpolarized $e^{+}$
beam allow to discriminate between scenario~A and scenario~B.
For $P_{-}=+85\%$, $P_{+}=0\%$ the FB--asymmetry is negative between
$A_{FB}=-8\%$ and $A_{FB}=-12\%$ in scenario~B, whereas it has large
positive values between $A_{FB}=31\%$ and $A_{FB}=34\%$ in scenario~A,
compare Fig.~6 and Fig.~8.
\section{Summary and Conclusions}
For the determination of the MSSM parameters $M$, $\mu$ and
$\tan\beta$ from chargino production, the sneutrino mass plays a
crucial role. With regard to the determination of the sneutrino mass we
have studied chargino pair production
$e^{+}e^{-}\to\tilde{\chi}^{+}_1\tilde{\chi}^{-}_1$ with polarized
beams and the subsequent leptonic decay
$\tilde{\chi}^{-}_1\to\tilde{\chi}^0_1 e^{-} \bar{\nu}_e$ taking into
account the complete spin correlations between production and decay.
The spin correlations are crucial for the decay angular
distributions and on the forward--backward asymmetries of
the decay leptons.
We have presented a method for constraining $m_{\tilde{\nu}_e}$
if $\tan\beta$ and the masses of $\tilde{\chi}^{-}_1$ and
of $\tilde{\chi}^0_1$ are known.
Accordingly we choose two representative scenarios with a dominating
gaugino and with a dominating higgsino component of
$\tilde{\chi}^{-}_1$, respectively.
We have studied the prospect to constrain
the sneutrino mass by measuring polarized cross sections and
FB--asymmetries of the decay leptons in the laboratory system.
Simultaneous polarization of both beams
is not only important for enlarging the cross section but it
can also have rather strong
influence on the FB--asymmetry of the decay
electron depending on
the mixing character of the charginos.
For charginos with dominating gaugino character the FB--asymmetry
is practically inpendent of the beam polarization.
But for higgsino-like charginos
beam polarization can considerably influence the FB--asymmetry.
In our examples $\sigma_{e^{-}}$ measured for
the beam polarization $P_{-}=-85\%$, $P_{+}=+60\%$ with
an error of 5\%, e.g., is
compatible with two regions for $m_{\tilde{\nu}_e}$.
For gaugino-like charginos as well as for higgsino-like ones
the corresponding FB--asymmetry allows to
distinguish between the two regions of $m_{\tilde{\nu}_e}$.
Beyond that the two mixing scenarios can be distinguished by
additional measurement of the FB--asymmetry with a right polarized
$e^{-}$ beam. Then the sign of $A_{FB}$ changes
for higgsino-like charginos.
We have presented numerical results for $\sqrt{s}=270$~GeV, near the
chargino production threshold, and for $\sqrt{s}=500$~GeV.
For gaugino-like charginos
$m_{\tilde{\nu}_e}$ can be constrained up to a few GeV, for
higgsino-like charginos this is possible if $m_{\tilde{\nu}_e}$ is in
the vicinity of the chargino mass.
Obviously a complete MC study with inclusion of experimental cuts
would be indispensable for realistic predictions.
\vspace{.4cm}
G.~M.--P.\ thanks M.~Jezabek and the other organizers of the Epiphany
Conference for the extremely hostly and
friendly atmosphere during the Conference.
We are grateful to V.~Latussek for his support in the development of
the numerical program. G.M.--P.\ was supported by {\it Stiftung f\"ur
Deutsch--Polnische Zusammenarbeit} and by {\it Friedrich--Ebert--Stiftung}.
\vspace{-.4cm}
|
\section{Introduction}
In this paper we study the variability and rigidity of secondary
characteristic classes which arise from flat connections $\theta$ on a
differentiable manifold $M$. In particular we consider $\theta$ as a
Lie-algebra valued one-form on $M$, and study the characteristic map
$$ \phi_{\theta}: H^*_{\text{Lie}}(\g ) \to H^*_{\text{dR}}(M), $$
where $H^*_{\text{Lie}}$ denotes Lie algebra cohomology (for a Lie
algebra $\g$), and $H^*_{\text{dR}}$ denotes de Rham cohomology. An
element $\alpha \in H^*_{\text{Lie}}(\g )$ is called variable if there exists
a one-parameter family of flat connections $\theta_t$ with
$$ \frac{d}{dt} \Big[ \phi_{\theta_t} (\alpha) \,
\Big{\vert}_{t=0} \neq 0, $$
otherwise $\alpha$ is called rigid. For example, the universal
Godbillon-Vey invariant for codimension $k$ foliations is known to be
a variable class. Letting $HL^*$ denote Loday's Leibniz
cohomology \cite{Ens.Math.}
\cite{overview} \cite{LP}, we prove that if $HL^n (\g )= 0$
for $n \geq 1$, then all classes in $H^*_{\text{Lie}}(\g )$ are rigid.
This result imparts a geometric meaning to Leibniz cohomology.
Moreover, in the case of codimension one foliations (with trivial
normal bundle), there is a flat $W_1$-valued connection on $M$, where
$W_1$ is the Lie algebra of formal vector fields on $\mathbf {R}^1$.
From \cite{Lodder},
$$ HL^*(W_1 ) \simeq \Lambda(\alpha) \otimes T(\zeta), $$
where $\Lambda (\alpha)$ is the exterior algebra on the Godbillon-Vey
invariant (in dimension three) and $T(\zeta)$ denotes the tensor
algebra on a four-dimensional class. When $M$ is provided with a
one-parameter family of such foliations, we compute the image of a
characteristic map
$$ HL^4 (W_1 ) \to H^4 _{\text{dR}}(M). $$
The image in de Rham cohomology is independent of the choices made
when constructing the $W_1$ connections, and involves the time
derivative (derivative with respect to $t$) of the Bott connection.
\section{The Characteristic Map and Rigidity}
Let $M$ be a differentiable ($C^{\infty}$) manifold with flat
connection $\theta$. We consider the formulation of $\theta$ as a
Lie-algebra valued one-form
$$ \theta : TM \to \g , $$
where $TM$ denotes the tangent bundle of $M$ and $\g$ is a Lie
algebra. This section describes a characteristic map
$$ \phi_{\theta} : H^*_{\text{Lie}} (\g ; \, \mathbf {R}) \to
H^*_{\text{dR}}(M) $$
from Lie algebra cohomology (with $\mathbf {R}$ coefficients) to the de Rham
cohomology of $M$. In the case of a topological Lie algebra, then
$H^*_{\text{Lie}}$ is understood as continuous cohomology, computed
using continuous cochains.
Let $\Omega^k (\g ; \, \mathbf {R} )$ be the $\mathbf {R}$-vector space of
skew-symmetric (continuous) cochains
$$ \alpha : \g^{\otimes k} \to \mathbf {R} . $$
For comparison with Leibniz cohomology (and to establish our sign
conventions), we write the coboundary for
Lie algebra cohomology
$$ d: \Omega^k (\g; \, \mathbf {R}) \to \Omega^{k+1}(\g; \, \mathbf {R}) $$
as
\begin{equation} \label{2.1}
\begin{split}
& d(\alpha)(g_1 \otimes g_2 \otimes \, \ldots \, \otimes g_{k+1}) = \\
& \sum_{1 \leq i < j \leq k} (-1)^{j+1} \alpha (g_1 \otimes \, \ldots \, g_{i-1}
\otimes [g_i, \, g_j] \otimes g_{i+1} \, \ldots \, \hat{g}_j \, \ldots \otimes g_{k+1}),
\end{split}
\end{equation}
where each $g_i \in \g$. For a differentiable manifold $M$, let
$$ \Omega^k (M) := \Omega^k(M; \, \mathbf {R}) $$
denote the $\mathbf {R}$-vector space of $k$-forms on $M$. Then the de Rham
coboundary
\begin{align*}
& d: \Omega^k (M) \to \Omega^{k+1}(M) \\
& \omega \mapsto d \omega
\end{align*}
has a global formulation as
\begin{equation} \label{2.2}
\begin{split}
& d \omega (X_1 \otimes X_2 \otimes \, \ldots \, \otimes X_{k+1}) = \\
& \sum_{i=1}^{k+1} (-1)^{i+1} X_i \big( \omega (X_1 \otimes \, \ldots \,
\hat{X}_i \, \ldots \otimes X_{k+1}) \big) + \\
& \sum_{1 \leq i < j \leq k+1} (-1)^{j+1} \omega \big( X_1 \otimes \, \ldots
\, \otimes X_{i-1}\otimes [X_i, \, X_j]\otimes X_{i+1}\otimes \, \\
& \hskip2.5in \ldots \,
\hat{X}_j \, \ldots \, \otimes X_{k+1} \big),
\end{split}
\end{equation}
where each $X_i \in \chi(M)$, the Lie algebra of smooth ($C^{\infty}$)
vector fields on $M$.
With the above sign conventions for the coboundary maps, the
connection $\theta \in \Omega^1(M; \, \g )$ is flat if and only if the
Maurer-Cartan equation holds,
\begin{equation} \label{2.3}
d \theta = - \frac{1}{2} \, [\theta, \, \theta],
\end{equation}
where $ d \theta \in \Omega^2 (M; \, \g)$ is given by
$$ (d \theta)(X_1 \otimes X_2) = X_1 (\theta(X_2)) - X_2(\theta (X_1))
- \theta([X_1, \, X_2]). $$
Recall that for a differentiable function $f : M \to \g$, $X(f)$
denotes the partial derivatives of the component functions of $f$
with respect to $X$. If $e_1$, $e_2$, $e_3$, $\ldots$ is a basis for
$\g$, then
$$ f = f_1 e_1 + f_2 e_2 + f_3 e_3 + \, \cdots , $$
where each $f_i : M \to \mathbf {R}$ is differentiable. Thus,
$$ X(f) = X(f_1)e_1 + X(f_2)e_2 + X(f_3)e_3 + \, \cdots .$$
Also, the symbol $[\theta, \, \theta]$ denotes the composition
$$ \Omega^1(M;\, \g) \otimes \Omega^1(M;\, \g)
\overset {\theta \wedge \theta}{\longrightarrow}
\Omega^2(M; \, \g \otimes \g)
\overset{[ \; , \; ]}{\longrightarrow} \Omega^2(M; \, \g)$$
given by
\begin{align*}
[\theta , \, \theta] (X_1 \otimes X_2)\, & = [ \ \, , \;] \circ \big(
\theta(X_1) \otimes \theta(X_2) - \theta(X_2) \otimes \theta(X_1)\big) \\
& = 2 [\theta (X_1), \, \theta(X_2)].
\end{align*}
Consider the following map \cite[p.\ 234]{Fuks}
$$ \phi: \Omega^*(\g ; \, \mathbf {R}) \to \Omega^*(M; \, \mathbf {R})$$
given on a cochain $\alpha \in \Omega^k(\g; \, \mathbf {R})$ by
$$
\phi(\alpha)(X_1 \otimes X_2 \otimes \, \ldots \, \otimes X_k) =
\alpha(\theta(X_1) \otimes \theta(X_2) \otimes \, \ldots \, \otimes \theta(X_k)).
$$
For the case $k=1$, it can be easily seen from (2.3) that $\phi$ is a
map of cochain complexes. In particular,
\begin{align*}
d(\phi(\alpha & )) (X_1 \otimes X_2) \\
& = X_1 (\alpha (\theta (X_2))) - X_2(\alpha (\theta (X_1))) -
\alpha(\theta( [X_1, \, X_2])) \\
& = \alpha \Big( X_1 (\theta(X_2)) - X_2(\theta(X_1)) -
\theta([X_1, \, X_2])\Big) \\
& = \alpha(- [\theta(X_1), \, \theta(X_2)]) \\
& = \phi(d\alpha)(X_1 \otimes X_2).
\end{align*}
In \cite[p.\ 235]{Fuks} it is proven that in general
$$ d \phi = \phi d . $$
The induced map
$$ \phi_{\theta} : H^*_{\text{Lie}}(\g ) \to H^*_{\text{dR}}(M) $$
is called the characteristic map.
A specific example of a flat connection studied in this paper arises
from the theory of foliations. Let $\cal F$ be a $C^{\infty}$
codimension one foliation on $M$ with trivial normal bundle. Given a
choice of a trivialization, then a determining one-form $\omega_0$ is
defined for the foliation by $\omega_0(v) = 0$ is $v$ is tangent to a
leaf and $\omega_0(\eta) = 1$ if $\eta$ is a unit vector of the normal
bundle with positive orientation. Letting $d$ denote the de Rham
coboundary, a sequence of one-forms $\omega_1$, $\omega_2$, $\omega_3$, $\ldots$
can be defined so that \cite{Godbillon}
\begin{equation} \label{2.4}
\begin{split}
& d \omega_0 = \omega_0 \wedge \omega_1 \\
& d \omega_1 = \omega_0 \wedge \omega_2 \\
& d \omega_2 = \omega_0 \wedge \omega_3 + \omega_1 \wedge \omega_2 \\
& d \omega_k = \sum_{i=0}^{\textstyle{[\frac{k}{2}}]} \, \, \frac{k-2i+1}{k+1}
\, \binom{k+1}{i} \, (\omega_i \wedge \omega_{k+1-i}).
\end{split}
\end{equation}
Consider the topological Lie algebra of formal vector fields
$$ W_1 = \Big\{ \, \sum_{k=0}^{\infty} c_k\, x^k \, \frac{d}{dx}
\ \ \Big{\vert}\ \ c_k \in \mathbf {R} \, \Big\} $$
in the $\cal M$-adic topology, where $\cal M$ is the maximal ideal of
$R[[x]]$ given by those series with zero constant term. Then a
$W_1$-valued one-form is defined on $M$ by
\begin{equation} \label{2.5}
\theta_{\cal F}(v) = \sum_{k=0}^{\infty} \omega_k(v) \,
\frac{x^k}{k!} \, \frac{d}{dx},
\end{equation}
where $v \in TM$. From (2.4), it can be proven \cite[p.\ 231]{Fuks}
that $\theta_{\cal F}$ is a flat connection on $M$. The resulting
homomorphism
\begin{equation} \label{1.6}
\phi_{\theta_{\cal F}}: H^*_{\text{Lie}}(W_1) \to H^*_{\text{dR}}(M)
\end{equation}
is the classical characteristic map in foliation theory. (A similar
construction exists for foliations of any codimension with trivial
normal bundle.)
We wish to study a one-parameter variation of a flat structure
\begin{equation} \label{2.7}
\theta_t : TM \to \g, \ \ \ t \in \mathbf {R}, \ \ \ \theta_0 = \theta,
\end{equation}
which depends smoothly on the parameter $t$. Such a structure may
arise from a one-parameter variation of a foliation.
\begin{definition} \label{2.8}
A Lie algebra cohomology class $\alpha \in H^*_{\text{Lie}}(\g)$ is called
variable (for $\theta$) if there exists a family $\theta_t$ such that
$$ \frac{d}{dt} \Big[ \phi_{\theta_t} (\alpha) \,
\Big{\vert}_{t=0} \neq 0. $$
Otherwise, $\alpha$ is called rigid.
\end{definition}
By work of Thurston, the universal Godbillon-Vey invariant $\alpha \in
H^3_{\text{Lie}}(W_1)$ is a variable class \cite{Thurston}. One of
the goals of
this paper is to prove that if the Leibniz cohomology of $\g$
vanishes, i.e., $HL^n(\g ) = 0$, $n \geq 1$, then all characteristic
classes in $H^*_{\text{Lie}}(\g )$ are rigid. In the remainder of
this section we restate the definition of rigidity in terms of a known
condition concerning $H^*_{\text{Lie}}(\g ;\, \g')$, the
Lie algebra cohomology of $\g$ with coefficients in the
coadjoint representation
$$ \g' = {\text{Hom}}^c_{\mathbf {R}}(\g, \, \mathbf {R}), $$
($c$ denotes continuous maps).
First introduce the current algebra $\tilde{\g} = C^{\infty}(\mathbf {R}, \,
\g )$ of differentiable maps from $\mathbf {R}$ to $\g$. Then given
$\theta_t$ as in (2.7), there is a flat $\tilde{\g}$ connection on $M$
\begin{equation} \label{2.9}
\begin{split}
& \Theta : TM \to \tilde{\g} \\
& \Theta(v)(t) = \theta_t(v), \ \ \ v \in TM,
\end{split}
\end{equation}
and a characteristic map
$$ \phi_{\Theta} : H^*_{\text{Lie}}( \tilde{\g}) \to H^*_{\text{dR}}(M).$$
Using an idea of D. Fuks \cite{Fuks},
define a ``time derivative'' map on cochains
$$ D: \Omega^q (\g) \to \Omega^q(\tilde{\g}) $$
by
\begin{align*}
& D(\alpha)(\varphi_1, \, \varphi_2, \, \ldots , \, \varphi_q) = \\
& \sum_{i=1}^{q} \alpha \Big(\varphi_1(0), \, \ldots , \,
\varphi_{i-1}(0), \, \frac{d}{dt}[\varphi_i(t)\,\big{\vert}_{t=0}\, , \,
\varphi_{i+1}(0), \, \ldots , \, \varphi_q (0) \Big),
\end{align*}
where $\alpha \in \Omega^q (\g )$, $(\varphi_1, \, \varphi_2, \, \ldots, \,
\varphi_q ) \in (\tilde{\g})^{\otimes q}$. Then $D$ is a map of cochain
complexes, and there is an induced map
\begin{equation} \label{1.10}
D^* : H^*_{\text{Lie}}(\g ) \to H^*_{\text{Lie}}(\tilde{\g}).
\end{equation}
It follows that given $\alpha \in H^*_{\text{Lie}}(\g )$, we have
\begin{equation} \label{2.11}
\phi_{\Theta} \circ D^* (\alpha) = \frac{d}{dt} \big[ \phi_{\theta_t}
(\alpha ) \big{\vert}_{t=0}\, .
\end{equation}
Recall that $\g' = \text{Hom}^c_{\mathbf {R}}(\g , \, \mathbf {R})$ is a left
$\g$-module with
$$ (g \gamma)(h) = \gamma([h, \, g]), $$
where $g, \, h \in \g$ and $\gamma \in \g'$. Then $D^*$ can be
factored as $\varPhi^* \circ V^*$ \cite[p.\ 244]{Fuks}, where
\begin{equation} \label{2.12}
\begin{split}
& V^* : H^q_{\text{Lie}}( \g) \to H^{q-1}(\g; \, \g'), \ \ q \geq 1, \\
& \varPhi^* : H^{q-1}(\g; \, \g') \to H^q_{\text{Lie}}(\tilde{\g}),
\ \ q \geq 1,
\end{split}
\end{equation}
are induced by
\begin{equation} \label{2.13}
\begin{split}
& \text{Var}: \Omega^q(\g; \, \mathbf {R}) \to \Omega^{q-1}(\g; \, \g') \\
& \varPhi : \Omega^{q-1}(\g; \, \g') \to \Omega^q(\tilde{\g}; \, \mathbf {R}) \\
& (\text{Var})(\alpha)(g_1, \, g_2, \, \ldots, \, g_{q-1})(g_0) =
(-1)^{q-1} \alpha(g_0, \, g_1, \, \ldots, \, g_{q-1}) \\
& \varPhi (\gamma)(\varphi_1, \, \varphi_2, \, \ldots, \, \varphi_q) = \\
& \ \ \ \ \
\sum_{i=1}^q (-1)^{q-i} \, \gamma \big( \varphi_1(0), \, \ldots, \,
\hat{\varphi}_i(0), \, \ldots, \, \varphi_q(0) \big) \big( \varphi'_i(0)\big),
\end{split}
\end{equation}
where $\alpha \in \Omega^q(\g; \, \mathbf {R})$, $g_i \in \g$,
$\gamma \in \Omega^{q-1}(\g; \, \g')$, $\varphi_i \in \tilde{\g}$.
We then have a commutative diagram
\begin{equation} \label{2.14}
\CD
H^*_{\text{Lie}}(\g) @>D^*>> H^*_{\text{Lie}}(\tilde{\g}) \\
@V{V^*}VV @AA{\varPhi^*}A \\
H^{*-1}_{\text{Lie}}(\g; \, \g') @>=>>
H^{*-1}_{\text{Lie}}(\g; \, \g')
\endCD
\end{equation}
\begin{lemma} \label{2.15}
If $H^{n-1}_{\text{Lie}}(\g; \, \g') = 0$ for $n \geq 1$, then all
characteristic classes in $H^*_{\text{Lie}}(\g )$ are rigid.
\end{lemma}
\begin{proof}
This follows from equation \eqref{2.11}, diagram \eqref{2.14} and the
definition of rigidity (definition \eqref{2.8}).
\end{proof}
In the next section we prove that if $HL^n(\g) = 0$ for $n \geq 1$,
then $H^{n-1}_{\text{Lie}}(\g; \, \g') = 0$ for $n \geq 1$.
\section{Leibniz Cohomology}
Still considering $\g$ to be a Lie algebra (over $\mathbf {R}$), recall that
the Leibniz cohomology of $\g$ with trivial coefficients,
$$ HL^*(\g; \, \mathbf {R}) := HL^*(\g ), $$
is the homology of the cochain complex \cite{LP}
\begin{equation} \label{3.1}
\mathbf {R} \overset{0}{\rightarrow} C^1(\g) \overset{d}{\rightarrow}
C^2(\g) \rightarrow \, \cdots \, \rightarrow C^k(\g) \overset{d}{\rightarrow}
C^{k+1}(\g) \rightarrow \, \cdots \, ,
\end{equation}
where $C^k(\g ) = {\text{Hom}}^c_{\mathbf {R}}( \g^{\otimes k}, \, \mathbf {R})$, and for
$\alpha \in C^k(\g)$, $d\alpha$ is given in equation \eqref{2.1}. Keep in mind that
for Leibniz cohomology, the cochains are not necessarily
skew-symmetric.
In this section we prove the following:
\begin{theorem} \label{3.2}
If $HL^n(\g; \, \mathbf {R}) = 0$ for $n \geq 1$, then
$H^{n-1}_{\text{Lie}}(\g; \, \g') = 0$ for $n \geq 1$, where $\g' =
{\text{Hom}}^c_{\mathbf {R}}(\g; \, \mathbf {R})$.
\end{theorem}
The proof involves a spectral sequence similar to the Pirashvili
spectral sequence \cite{Pirashvili}, except tailored to the specific
algebraic relation between $HL^*(\g)$ and $H^{*-1}_{\text{Lie}}(\g; \,
\g')$. Recall that the projection to the exterior power
$$ \g^{\otimes q} \to \g^{\wedge q} $$
induces a homomorphism
$$ H^*_{\text{Lie}}(\g) \to HL^*(\g). $$
Letting $C^*_{\text{rel}}(\g)[2] = C^*(\g)/ \Omega^*(\g)$,
we have a long exact sequence
$$ \cdots \, \rightarrow H^q_{\text{Lie}}(\g ) \rightarrow HL^q(\g)
\rightarrow H^{q-2}_{\text{rel}}(\g) \rightarrow
H^{q+1}_{\text{Lie}}(\g ) \rightarrow \, \cdots \, . $$
The Pirashvili spectral sequence arises from a filtration of
$C^*_{\text{rel}}(\g)[2]$ and converges to $H^*_{\text{rel}}(\g)$.
Consider now the map of cochain complexes
$$ i: \Omega^{q-1}(\g; \, \g') \to C^q({\g}) $$
given by
$$ \big( i(\beta )\big) (g_0 \otimes g_1 \otimes \, \ldots \, \otimes g_{q-1})
= (-1)^{q-1} \, \beta(g_1 \otimes g_2 \otimes \, \ldots \, \otimes g_{q-1})(g_0), $$
where $\beta \in \Omega^{q-1}(\g; \, \g')$ and $g_i \in \g$ for
$i = 0, \, 1, \, 2, \, \ldots, \, q-1$. Letting
$$ C^*_{RG}(\g)[2] = C^*(\g)/i[\Omega^{*-1}(\g; \, \g')], $$
we also have a long exact sequence
\begin{equation} \label{3.3}
\begin{split}
& 0 \rightarrow H^0_{\text{Lie}}(\g; \, \g') \rightarrow HL^1(\g)
\rightarrow 0 \rightarrow \\
& H^1_{\text{Lie}}(\g; \, \g') \rightarrow
HL^2(\g) \rightarrow H^0_{RG}(\g) \rightarrow
H^2_{\text{Lie}}(\g; \, \g')\rightarrow \, \\
& \cdots \, \rightarrow H^{q-1}_{\text{Lie}}(\g; \, \g') \rightarrow
HL^q(\g) \rightarrow H^{q-2}_{RG}(\g) \rightarrow H^q_{\text{Lie}}(\g; \, \g')
\rightarrow \, \cdots \, .
\end{split}
\end{equation}
The filtration for the Pirashvili spectral sequence \cite{Lodder}
\cite{Pirashvili} can be immediately applied to yield a decreasing
filtration $\{ F^s_* \}_{s \geq 0}$ for $C^*_{RG}(\g)[2]$. We use the
same grading as in \cite{Lodder} \cite{Pirashvili}, which becomes
$F^0_* = C^*_{RG}[2]$, and for $s \geq 1$,
\begin{align*}
& F^s_* = A/B \\
& A = \{ \, f \in C^*(\g) \ | \ f \ {\text{is skew-symmetric in the
last $(s+1)$ tensor factors}} \, \} \\
& B = i [ \Omega^{*-1}(\g; \, \g')].
\end{align*}
Then as in \cite{Lodder}, each $F^s_*$ is a subcomplex of
$C^*_{RG}(\g)$, and
$$ F^0_* \supseteq F^1_* \supseteq F^2_*
\supseteq \, \ldots \, \supseteq F^s_* \supseteq F^{s+1}_*
\supseteq \, \ldots \, . $$
To identify the $E^2$ term of the resulting spectral sequence,
consider coker(Var), where Var is defined in equation \eqref{2.13}.
Letting
$$ CR^*(\g)[1] = \Omega^{*-1}(\g; \, \g')/{\text{Var}}[\Omega^*(\g)], $$
there is a short exact sequence
\begin{equation} \label{3.4}
0 \rightarrow \Omega^*(\g) \overset{\text{Var}}{\longrightarrow}
\Omega^{*-1}(\g; \, \g') \rightarrow CR^*(\g)[1] \rightarrow 0,
\end{equation}
and an associated long exact sequence
$$ \cdots \, \rightarrow H^q_{\text{Lie}}(\g) \rightarrow
H^{q-1}_{\text{Lie}}(\g; \g') \rightarrow HR^{q-2}(\g) \rightarrow
H^{q+1}_{\text{Lie}}(\g ) \rightarrow \, \cdots \, . $$
\begin{theorem} \label{3.5}
The filtration $F^s_*$ of $C^*_{RG}(\g)[2]$ yields a spectral sequence
converging to $H^*_{RG}(\g)$ with
\begin{align*}
& E^{s,0}_2 = 0, \ \ s = 0, \, 1, \, 2, \, \ldots, \\
& E^{s,n}_2 \simeq HL^n(\g) \hat{\otimes} HR^s(\g), \ \ n = 1, \, 2, \, 3,
\, \ldots, \ \ s = 0, \, 1, \, 2, \, \ldots ,
\end{align*}
where $\hat{\otimes}$ denotes the completed tensor product.
\end{theorem}
\begin{proof}
The proof follows from the identification of the $E^2$ term in
\cite{Lodder} or \cite{Pirashvili}. Also, note that
$$ E^{s,0}_0 = (F^s/F^{s+1})_0, \ \ \ F^s_0 = A/B, $$
where
\begin{align*}
& A = \{ \, f \in C^{s+2}(\g) \ | \ f \ {\text{is alternating in the
last $(s+1)$ factors}} \, \} \\
& B = i[\Omega^{s+1}(\g; \g')].
\end{align*}
Then $A = B$, $F^s_0 = 0$, and $E^{s,0}_0 = 0$.
\end{proof}
\begin{theorem} \label{3.6}
If $HL^n(\g) = 0$ for $n \geq 1$, then $H^{n-1}_{\text{Lie}}(\g; \,
\g') = 0$ for $n \geq 1$.
\end{theorem}
\begin{proof}
If $HL^n(\g) = 0$ for $n \geq 1$, then from theorem \eqref{3.5}, the
$E_2$ term for the spectral sequence converging to $H^*_{RG}(\g)$ is
zero. Thus, $H^n_{RG}(\g) = 0$ for $n \geq 0$. The result now
follows from long exact sequence \eqref{3.3}.
\end{proof}
\begin{theorem} \label{3.7}
If $HL^n(\g) = 0$ for $n \geq 1$, and $\theta_t$ is a one-parameter
family of flat $\g$-connections on $M$, then all characteristic
classes in $H^*_{\text{Lie}}(\g)$ are rigid.
\end{theorem}
\begin{proof}
The theorem follows from lemma \eqref{2.15} and theorem \eqref{3.6}.
\end{proof}
By checking dimensions in theorem \eqref{3.5}, exact sequence
\eqref{3.3}, and diagram \eqref{2.14}, we have:
\begin{corollary} \label{3.8}
If $HL^n(\g) = 0$ for $1 \leq n \leq p$, then all characteristic
classes in $H^n_{\text{Lie}}(\g)$ are rigid for $1 \leq n \leq p$.
\end{corollary}
We close this section with two observations, one concerning a theorem
of P. Ntolo on the vanishing of $HL^* (\g)$ for $\g$ semi-simple, the
other concerning the highly nontrivial nature of $HL^* (W_1)$.
\begin{theorem} \label{3.9}
\cite{Ntolo} If $\g$ is a semi-simple Lie algebra (over $\mathbf {R}$), then
$$ HL^n(\g) = 0 \ \ \ {\text{for}} \ \, n \geq 1. $$
\end{theorem}
By contrast, the Leibniz cohomology of formal vector fields,
$HL^*(W_1)$, contains many non-zero classes which do not appear in
$H^*_{\text{Lie}}(W_1)$ \cite{Lodder}. In the next section we compute
the image of a characteristic map
$$ HL^4(W_1) \to H^4_{\text{dR}}(M), $$
where $M$ supports a family of codimension one foliations.
\section{Foliations}
Letting $W_1$ denote the Lie algebra of formal vector fields defined
in section two, recall that \cite[p.\ 101]{Fuks}
$$ H^{q-1}_{\text{Lie}}(W_1; \, W'_1) \simeq \mathbf {R} $$
for $q=3$ and $q=4$, and zero otherwise.
The generator of the class for $q=3$ is the
universal Godbillon-Vey invariant, called $\alpha$ in this paper, and we
denote the generator of the class for $q=4$ by $\zeta$. From
\cite{Lodder}, the map
$$ H^{*-1}_{\text{Lie}}(W_1; \, W'_1) \to HL^*(W_1) $$
given in exact sequence \eqref{3.3} is injective. As dual Leibniz
algebras \cite{Bourbaki}, we have \cite{Lodder}
\begin{equation} \label{4.1}
HL^*(W_1 ) \simeq \Lambda(\alpha) \otimes T(\zeta),
\end{equation}
where $\Lambda (\alpha)$ is the exterior algebra on $\alpha$, and $T(\zeta)$
denotes the tensor algebra on $\zeta$.
Let $M$ be a $C^{\infty}$ manifold with a one-parameter family ${\cal
F}_t$ of codimension one foliations having trivial normal bundles.
Let $\omega_i(t)$ be the corresponding one-forms given in equation
\eqref{2.4} considered as differentiable functions of $t$. Recall the
definitions of $\varPhi^*$ and $\phi_{\Theta}$ given in equations \eqref{2.12}
and \eqref{2.9} respectively.
In this section we prove the following:
\begin{theorem} \label{4.2}
Let $M$ and ${\cal F}_t$ be given as above. Then the composition
$$ HL^4(W_1) \simeq H^3_{\text{Lie}}(W_1; \, W'_1)
\overset{\varPhi^*}{\rightarrow} H^4_{\text{Lie}}(\tilde{W}_1)
\overset{\phi_{\Theta}}{\rightarrow} H^4_{\text{dR}}(M) $$
sends $\zeta$ to the de Rham cohomology class of
$$ c(\zeta):= \omega'_1(0) \wedge \omega_0(0) \wedge \omega_1(0) \wedge \omega_2(0).$$
Moreover, the cohomology class of $c(\zeta)$ does not depend on the
choice\footnote{Since $c(\zeta)$ may also be written as
$- \omega'_1(0) \wedge \omega_1(0) \wedge d\omega_1(0)$, where $d$ denotes the
de Rham coboundary, it is not necessary to show that the class of $c(\zeta)$ is
independent of the choice of $\omega_2(t)$.}
of $\omega_0(t)$ or $\omega_1(t)$.
\end{theorem}
\begin{proof}
We first compute $c(\zeta)$ on the level of cochains. Consider the
vector space basis $\{ \beta_i \}_{i \geq 0}$ of
${\text{Hom}}^c_{\mathbf {R}}(W_1, \, \mathbf {R})$ given by
$$ \beta_i \Big( \frac{x^j}{j!} \, \frac{d}{dx}\Big) = \delta_{ij}. $$
From \cite{Lodder}, the class of $\zeta$ in $HL^4 (W_1)$ is
represented by the cochain
$$ \beta_1 \otimes (\beta_0 \wedge \beta_1 \wedge \beta_2). $$
The cochain map
$$ i: \Omega^3(W_1; \, W'_1) \to C^4(W_1) $$
inducing the isomorphism
$$ H^3_{\text{Lie}}(W_1; \, W'_1) \overset{\simeq}{\rightarrow}
HL^4(W_1) $$
satisfies
$$ i\big( ( \beta_0 \wedge \beta_1 \wedge \beta_2) \otimes \beta_1 \big) =
- \beta_1 \otimes ( \beta_0 \wedge \beta_1 \wedge \beta_2). $$
Also, it is know that
$$ ( \beta_0 \wedge \beta_1 \wedge \beta_2) \otimes \beta_1 $$
generates $H^3_{\text{Lie}}(W_1; \, W'_1)$ (as an $\mathbf {R}$ vector space).
Let $v_1$, $v_2$, $v_3$, $v_4 \ \in T_p(M)$. From the definition of
$\varPhi$ and $\phi_{\Theta}$, the image of $\zeta$ in $\Omega^4(M)$
is the 4-form which sends $v_1 \otimes v_2 \otimes v_3 \otimes v_4$ to
\begin{align*}
- ( \beta_1 \otimes \beta_0 \wedge \beta_1 \wedge \beta_2) \Big( & -A_1' \otimes A_2
\otimes A_3 \otimes A_4 + A'_2 \otimes A_1 \otimes A_3 \otimes A_4 \\
& -A'_3 \otimes A_1 \otimes A_2 \otimes A_4 + A'_4 \otimes A_1 \otimes A_2 \otimes A_3\Big),
\end{align*}
where
\begin{align*}
& A_i = \sum_{n \geq 0} \, \omega_n(0)(v_i) \, \frac{x^n}{n!} \,
\frac{d}{dx} \, , \ \ \ i = 1, \ 2, \ 3, \ 4, \\
& A'_i = \sum_{n \geq 0} \, \omega'_n(0)(v_i) \, \frac{x^n}{n!} \,
\frac{d}{dx} \, , \ \ \ i = 1, \ 2, \ 3, \ 4.
\end{align*}
By the definition of the $\beta_i$'s, the image of $\zeta$ is thus
$$ \big( \omega'_1(0) \wedge \omega_0(0) \wedge \omega_1(0) \wedge \omega_2(0)\big)
(v_1 \otimes v_2 \otimes v_3 \otimes v_4). $$
To show that the de Rham cohomology class of $c(\zeta)$ does not
depend on the choice of $\omega_0(t)$, consider the one-form
$$ u_0(t) = f \cdot \omega_0(t), $$
where $f:M \to \mathbf {R}$ is a $C^{\infty}$ function with $f(p) \neq 0$ for
all $p \in M$. Letting $d$ denote the de Rham coboundary, we have
from equation \eqref{2.4}
$$ d \omega_1(t) = \omega_0(t) \wedge \omega_2(t). $$
Then
$$ \omega'_0(0) \wedge \omega_0(0) \wedge \omega_1(0) \wedge \omega_2(0) =
- \omega'_0(0) \wedge \omega_1(0) \wedge d \omega_1(0). $$
Also,
\begin{align*}
& d u_0(t) = u_0(t) \wedge \Big( \frac{-df}{f} + \omega_1(t) \Big) \\
& u_1(t) = \frac{-df}{f} + \omega_1(t) \\
& u'_1(t) \wedge u_1(t) \wedge du_1(t) = \omega'_1(t) \wedge \omega_1(t)
\wedge d \omega_1(t) \\
& \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
+ \frac{df}{f} \wedge \omega'_1(t) \wedge d \omega_1(t).
\end{align*}
It follows that
\begin{align*}
u'_1(0) \wedge u_1(0) \wedge du_1(0) = \; & \omega'_1(0) \wedge \omega_1(0)
\wedge d \omega_1(0) \\
& + d\big( \log ( |f| ) \, \omega'_1(0) \wedge d \omega_1(0) \big).
\end{align*}
Compare with Ghys \cite{Ghys}. Of course,
$$ d(\omega'_1(0)) = \omega'_0(0) \wedge \omega_2(0) + \omega_0(0) \wedge
\omega'_2(0). $$
To show that the cohomology class of $c(\zeta)$ does not depend on the
choice of $\omega_1(t)$, consider the one-forms
$$ u(t) = \omega_1(t) + f \cdot \omega_0(t), $$
where $ g: M \to \mathbf {R}$ is a $C^{\infty}$ function (which may have
zeroes on $M$). Then
\begin{align*}
u'(0) \wedge u(0) \wedge du(0)& = \omega'_1(0) \wedge \omega_1(0) \wedge d\omega_1(0)\\
& + g\cdot \omega'_0(0) \wedge \omega_1(0) \wedge d\omega_1(0) \\
& + \omega'_1(0) \wedge \omega_1(0) \wedge dg \wedge \omega_0(0) \\
& + g \cdot \omega'_0(0) \wedge \omega_1(0) \wedge dg \wedge \omega_0(0).
\end{align*}
It can be checked that
\begin{align*}
& u'(0) \wedge u(0) \wedge du(0) = \omega'_1(0) \wedge \omega_1(0) \wedge
d\omega_1(0) + d(A), \\
& A = g \cdot \omega'_0(0) \wedge d\omega_1(0) - dg \wedge \omega'_0(0) \wedge
\omega_1(0) - \frac{1}{2} \, g^2 \cdot \omega'_0(0) \wedge d \omega_0(0).
\end{align*}
\end{proof}
The paper is closed by noting that the current algebra $\tilde{\g}$ is
a Leibniz algebra in the sense of Loday \cite{Ens.Math.} with the
Leibniz bracket of $\varphi_1, \; \varphi_2 \in \tilde{\g}$ given by
$$ \langle \varphi_1(t), \, \varphi_2(t) \rangle =
[\varphi_1(t), \, \varphi'_2 (0)]_{\text{Lie}}, $$
where $[ \ \, , \ ]_{\text{Lie}}$ is the usual Lie bracket on
$\tilde{\g}$, and $\varphi'_2(0)$ is the constant path at
$\varphi'_2(0)$. The Leibniz bracket is
not necessarily skew-symmetric,
$$ \langle \varphi_1(t), \, \varphi_2(t) \rangle \neq
- \langle \varphi_2(t), \, \varphi_1(t) \rangle , $$
but satisfies the following version of the Jacobi identity
$$ \langle \varphi_1(t), \, \langle \varphi_2(t), \, \varphi_3(t) \rangle \rangle
= \langle \langle \varphi_1(t), \, \varphi_2(t) \rangle, \,
\varphi_3(t) \rangle - \langle \langle \varphi_1(t) , \,
\varphi_3(t) \rangle, \, \varphi_2(t) \rangle, $$
which is the defining relation for a Leibniz algebra. Also see
\cite{Balavoine} and \cite{Ens.Math.}.
\bigskip
\noindent
ACKNOWLEDGEMENTS
\smallskip
\noindent
The author would like to express his gratitude to the Institut des
Hautes \'Etudes Scientifiques for their support while this paper was being
written.
|
\section{Introduction}
It is generally assumed that our Universe contains an approximately
equal amount of leptons and antileptons. The lepton asymmetry would be
of the same order as the baryon asymmetry, which is very small as
required by Big Bang Nucleosynthesis (BBN) considerations. The
existence of a large lepton asymmetry is restricted to be in the form
of neutrinos from the requirement of universal electric neutrality,
and the possibility of a large neutrino asymmetry is still open. {}From
a particle physics point of view, a lepton asymmetry can be generated
by an Affleck-Dine mechanism \cite{AF} without producing a large
baryon asymmetry (see ref.~\cite{Casas} for a recent model), or even
by active-sterile neutrino oscillations after the electroweak phase
transition (but in this last case, it might not be of order unity)
\cite{Foot}. This lepton asymmetry can postpone symmetry restoration
in non-supersymmetric or supersymmetric models \cite{Linde} (note
that this is also true for other charges \cite{Riotto}).
In this paper we study some cosmological implications of relic
degenerate neutrinos\footnote{Here, by degeneracy, we mean that there exists a
large neutrino-antineutrino asymmetry, or vice versa, and not a
degeneracy in the mass sense.}. We do not consider any specific model
for generating such an asymmetry, and just assume that it was created
well before neutrinos decouple from the rest of the plasma. An
asymmetry of order one or larger can have crucial effects on the
global evolution of the Universe. Among other effects, it changes the
decoupling temperature of neutrinos, the primordial production of
light elements at BBN, the time of equality between radiation and
matter, or the contribution of relic neutrinos to the present energy
density of the Universe. The latter changes affect the evolution of
perturbations in the Universe. We focus on the anisotropies of the
Cosmic Microwave Background (CMB), and on the distribution of Large
Scale Structure (LSS). We calculate the power spectrum of both
quantities, in the case of massless degenerate neutrinos, and also for
neutrinos with a mass of $0.07$ eV, as suggested to explain the
experimental evidence of atmospheric neutrino oscillations at
Super-Kamiokande \cite{SK}. The cosmological implications of
neutrinos with such a small mass are known to be very small,
but we will see that this conclusion is modified if a large
neutrino degeneracy exists. We also include in our analysis the
possibility that the dominant contribution to the present energy
density in the Universe is due to a cosmological constant:
$\Omega_\Lambda \sim 0.6-0.7$, keeping the Universe flat
($\Omega_0+\Omega_\Lambda=1$), as suggested by recent observations
(see \cite{Lineweaver} and references therein).
The effect of neutrino degeneracy on the LSS power spectrum was
studied in ref.~\cite{Larsen}, as a way of improving the agreement
with observations of mixed dark matter models with eV neutrinos, in
the case of high values of the Hubble parameter. Also, Adams \&
Sarkar \cite{Sarkar} calculated the CMB anisotropies and the matter
power spectrum, and compared them with observations in the
$\Omega_\Lambda=0$ case for massless degenerate neutrinos. More
recently, Kinney \& Riotto \cite{Kinney} also calculated the CMB
anisotropies for massless degenerate neutrinos in the
$\Omega_\Lambda=0.7$ case.
This paper is organized as follows. In section \ref{energy}, we
calculate the contribution of massive degenerate neutrinos to the
present energy density of the Universe. In section \ref{power}, we
explain how to calculate the power spectra, with the help of the code
{\tt cmbfast} \cite{SelZal}. In section \ref{results}, we discuss the
effect of the degeneracy on CMB anisotropies and the matter power
spectrum, both for massless neutrinos and $m_{\nu}=0.07$ eV. Finally,
in section \ref{comparison}, we derive some constraints on the
neutrino degeneracy from CMB and LSS data, in the particular case of a
flat Universe with an arbitrary cosmological constant and for standard
values of other cosmological parameters.
\section{Energy density of massive degenerate neutrinos}
\label{energy}
The energy density of one species of massive degenerate neutrinos and
antineutrinos, described by the distribution functions $f_\nu$ and
$f_{\bar{\nu}}$, is (here and throughout the paper we use
$\hbar=c=k_B=1$ units)
\begin{equation}
\rho_\nu + \rho_{\bar{\nu}}=
\int \frac{d^3\vec{p}}{(2\pi)^3} ~E_\nu (f_\nu(p)+ f_{\bar{\nu}}(p))=
\frac{1}{2\pi^2}\int_0^\infty dp ~p^2 \sqrt{p^2 + m_\nu^2}
(f_\nu(p)+ f_{\bar{\nu}}(p))~,
\label{defrhonu}
\end{equation}
valid at any moment. Here $p$ is the magnitude of the 3-momentum and
$m_\nu$ is the neutrino mass.
When the early Universe was hot enough, the neutrinos were in
equilibrium with the rest of the plasma via the weak interactions. In
that case the distribution functions $f_\nu$ and $f_{\bar{\nu}}$
changed with the Universe expansion, keeping the form of a Fermi-Dirac
distribution,
\begin{equation}
f_\nu(p)=\Frac{1}{\exp \left(\frac{p}{T_\nu} -
\frac{\mu}{T_\nu}\right)+1} \qquad
f_{\bar{\nu}}(p)=\Frac{1}{\exp
\left(\frac{p}{T_\nu} + \frac{\mu}{T_\nu}\right)+1}
\label{FD}
\end{equation}
Here $\mu$ is the neutrino chemical potential, which is nonzero if a
neutrino-antineutrino asymmetry has been previously produced. Later
the neutrinos decoupled when they were still
relativistic\footnote{Unless the neutrino mass is comparable to the
decoupling temperature, ${\cal O}(m_\nu) \sim$ MeV.}, and from that
moment the neutrino momenta just changed according to the cosmological
redshift. If $a$ is the expansion factor of the Universe, the
neutrino momentum decreases keeping $ap$ constant. At the same time
the neutrino degeneracy parameter $\xi \equiv \mu/T_\nu$ is conserved,
with a value equal to that at the moment of decoupling. Therefore one
can still calculate the energy density of neutrinos now from
\eq{defrhonu} and \eq{FD}, replacing $\mu/T_\nu$ by $\xi$ and
$p/T_\nu$ by $p/(y_\nu T_0)$, where $T_0 \simeq 2.726$ K and $y_\nu$
is the present ratio of neutrino and photon temperatures, which is not
unity because once decoupled the neutrinos did not share the entropy
transfer to photons from the successive particle annihilations that
occurred in the early Universe. In the standard case, the massless
non-degenerate neutrinos decoupled just before the electron-positron
pairs annihilated to photons, from which one gets the standard factor
$y_\nu=(4/11)^{1/3}$.
In the presence of a significant neutrino degeneracy $\xi$ the
decoupling temperature $T(\xi)$ is higher than in the standard
case \cite{Freese,Kang}. The reaction rate $\Gamma$ of the weak
processes, that keep the neutrinos in equilibrium with the other
species, is reduced because some of the initial or final neutrino
states will be occupied. The authors of ref.~\cite{Kang} used the
Boltzmann equation to calculate $\Gamma$ for the process $\nu_d +
\bar{\nu}_d \leftrightarrow e^+ + e^-$ (here $\nu_d$ denotes
degenerate neutrinos), including the corresponding Fermi blocking
factors. It was found that the neutrino decoupling temperature is
$T_{dec}(\xi) \approx 0.2\xi^{2/3}\exp(\xi/3)$ MeV (for
$\nu_\mu$ or $\nu_\tau$). Therefore if $\xi$ is large enough,
the degenerate neutrinos decouple before the temperature of the
Universe drops below the different mass thresholds, and are not
heated by the particle-antiparticle annihilations. The ratio of
neutrino and photon temperatures is thus reduced accordingly.
The present contribution of these degenerate neutrinos to the energy
density of the Universe can be parametrized as $\rho_\nu = 10^4 h^2
\Omega_\nu$ eV cm$^{-3}$, where $\Omega_\nu$ is the neutrino energy
density in units of the critical density $\rho_c=3H^2M_P^2/8\pi$,
$M_P=1.22 \times 10^{19}$ GeV is the Planck mass and $H=100h$ Km
s$^{-1}$ Mpc$^{-1}$ is the Hubble parameter.
The value of $\rho_\nu$ can be calculated as a function of the
neutrino mass and the neutrino degeneracy $\xi$, or equivalently the
present neutrino asymmetry $L_\nu$ defined as the following ratio of
number densities
\begin{equation}
L_\nu \equiv \frac{n_\nu-n_{\bar{\nu}}}{n_\gamma} =
\frac{1}{12\zeta (3)} y^3_\nu [\xi^3 + \pi^2 \xi]
\label{Lnu}
\end{equation}
We show\footnote{Here we assume $\xi>0$, but the results are also
valid for $\xi<0$ provided that $\xi$ and $L_\nu$ are understood as
moduli.} in figures \ref{ximass} and \ref{lnumass} the contours in the
$(m_\nu,L_\nu)$ and $(m_\nu,\xi)$ planes that correspond to some
particular values of $h^2 \Omega_\nu$. One can see from the figures
that there are two limits: massive non-degenerate neutrinos and
massless degenerate neutrinos. The first case corresponds to the
vertical lines, when one recovers the well-known bound on the neutrino
mass $m_\nu \raise0.3ex\hbox{$\;<$\kern-0.75em\raise-1.1ex\hbox{$\sim\;$}} 46$ eV for $h^2 \Omega_\nu=0.5$. On the other hand,
for very light neutrinos, the horizontal lines set a maximum value on the
neutrino degeneracy, that would correspond to a present neutrino
chemical potential $\mu_0 \raise0.3ex\hbox{$\;<$\kern-0.75em\raise-1.1ex\hbox{$\sim\;$}} 7.4 \times 10^{-3}$ eV, also for $h^2
\Omega_\nu=0.5$. In the intermediate region of the figures the
neutrino energy density is $\rho_\nu \simeq m_\nu n_\nu (\xi)$ and the
contours follow roughly the relations
$$
L_\nu \left (\frac{m_\nu}{\mbox{eV}} \right )\simeq 24.2 h^2\Omega_\nu
$$
\begin{equation}
(\pi^2\xi + \xi^3) \left (\frac{m_\nu}{\mbox{eV}} \right )
\simeq \frac{350}{y_\nu^3} h^2\Omega_\nu
\label{lines2}
\end{equation}
A similar calculation has been recently performed in reference
\cite{PalKar}. However the difference between neutrino and photon
temperatures was not properly taken into account for large $\xi$.
It was argued that, since the number density of highly degenerate neutrinos is
larger than in the non-degenerate case, the neutrinos would have been longer in
thermal contact with $e^+e^-$, therefore sharing with photons the entropy release.
However this is not the case \cite{Kang} as we discussed before.
The presence of a neutrino degeneracy can modify the outcome of Big
Bang Nucleosynthesis (for a review see \cite{Sarkar96}). First a
larger neutrino energy density increases the expansion rate of the
Universe, thus enhancing the primordial abundance of $^4$He. This is
valid for a nonzero $\xi$ of any neutrino flavor. In addition if the
degenerate neutrinos are of electron type, they have a direct
influence over the weak processes that interconvert neutrons and
protons. This last effect depends on the sign of $\xi_{\nu_e}$. Both
effects may be simultaneously important and it could be possible in
principle to explain the observed primordial abundances with a large
baryon density, $\Omega_B h^2 \approx 1$ \cite{Freese,Kang}. However
this possibility is ruled out by the fact that in that case our
Universe would have been radiation dominated during a longer period
and the observed large-scale structure would be difficult to
explain. {}From BBN one gets the following constraint \cite{Kang}
\begin{equation}
-0.06 \raise0.3ex\hbox{$\;<$\kern-0.75em\raise-1.1ex\hbox{$\sim\;$}} \xi_{\nu_e} \raise0.3ex\hbox{$\;<$\kern-0.75em\raise-1.1ex\hbox{$\sim\;$}} 1.1
\label{bbn}
\end{equation}
while a sufficiently long matter dominated epoch requires
\begin{equation}
|\xi_{\nu_\mu,\nu_\tau}| \raise0.3ex\hbox{$\;<$\kern-0.75em\raise-1.1ex\hbox{$\sim\;$}} 6.9
\label{lsf}
\end{equation}
This estimate from \cite{Kang} agrees with our analysis in section
\ref{comparison}. Assuming that the degenerate neutrinos are
$\nu_\mu$ or $\nu_\tau$, this places a limit on the degeneracy as shown by
the horizontal line in figures \ref{ximass} and \ref{lnumass}.
\section{Power spectra calculation}
\label{power}
We compute the power spectra of CMB anisotropies and large-scale
structure using the Boltzmann code {\tt cmbfast} by Seljak \&
Zaldarriaga \cite{SelZal}, adapted to the case of one family of
degenerate neutrinos ($\nu$, $\bar{\nu}$), with mass $m_\nu$ and
degeneracy parameter $\xi$. Let us first review the required
modifications. We use the notations of Ma \& Bertschinger
\cite{MaBer}, and for all issues not specific to our case, we refer
the reader to this review.
Background quantities can be rewritten in terms of two dimensionless
parameters ($M$, $Q$)
\begin{equation} \label{M}
M = \frac{m_\nu}{T_{\nu 0}}=
\frac{m_\nu(\mbox{eV})}{8.6170 \times 10^{-5}\times(4/11)^{1/3}~T_0(\mbox{K})},
\qquad
Q = \frac{a~p}{T_{\nu 0}}
\end{equation}
(we are assuming $y_\nu = (4/11)^{1/3}$, and therefore $\xi \leq 12$ \cite{Kang};
the scale factor is defined so that $a = 1$ today). For
Super-Kamiokande neutrinos with $m_\nu=0.07$ eV, $M \simeq 417$. We
then get for the mean density, pressure and phase-space distributions
\begin{eqnarray} \label{background}
\bar{\rho}_\nu + \bar{\rho}_{\bar{\nu}}
&=&
\frac{T^4_{\nu}}{2 \pi^2}
\int Q^2 dQ \sqrt{Q^2 + a^2 M^2}
\left( f_{\nu} (Q) + f_{\bar{\nu}} (Q) \right),
\nonumber \\
\bar{P}_\nu + \bar{P}_{\bar{\nu}}
&=&
\frac{T^4_{\nu}}{6 \pi^2}
\int Q^2 dQ
\frac{Q^2}{\sqrt{Q^2 + a^2 M^2}}
\left( f_{\nu} (Q) + f_{\bar{\nu}} (Q) \right),
\\
f_{\nu}(Q) &=& \frac{1}{e^{Q-\xi}+1}, \qquad
f_{\bar{\nu}}(Q) = \frac{1}{e^{Q+\xi}+1}. \nonumber
\end{eqnarray}
In the case of massive degenerate neutrinos, these integrals must be
calculated for each value of the scale factor, and also at the
beginning of the code in order to find $\Omega_{\nu}$ today. On the
other hand, for massless neutrinos, there is an exact analytic
solution
\begin{equation}
\bar{\rho}_{\nu} + \bar{\rho}_{\bar{\nu}}
= 3 (\bar{P}_{\nu} + \bar{P}_{\bar{\nu}} )
= \frac{7}{8}
\frac{\pi^2}{15}
T^4_{\nu}
\left[ 1+
\frac{30}{7}
\left(
\frac{\xi}{\pi}
\right)^2 + \frac{15}{7} \left( \frac{\xi}{\pi} \right)^4
\right].
\end{equation}
So, if we define an effective number of massless neutrino families
$N_{eff} \equiv 3 + 30/7 (\xi / \pi)^2 + 15/7 ( \xi / \pi )^4$,
the mean density and pressure for all neutrinos will be given by these
ones for one massless non-degenerate family, multiplied by $N_{eff}$.
Let us now consider perturbed quantities. We define $\Psi_{\nu}$ and
$\Psi_{\bar{\nu}}$, the perturbations of the phase space distribution for
$\nu$ and $\bar{\nu}$, through
\begin{eqnarray}
\delta f_{\nu} (\vec{x}, Q, \hat{n}, \tau) &=& f_{\nu} (Q) \Psi_{\nu} (\vec{x}, Q, \hat{n}, \tau)~, \
\nonumber \\
\delta f_{\bar{\nu}} (\vec{x}, Q, \hat{n}, \tau) &=& f_{\bar{\nu}} (Q) \Psi_{\bar{\nu}} (\vec{x}, Q,
\hat{n}, \tau)
\end{eqnarray}
($\hat{n}$ is the momentum direction: $\vec{p} \equiv p \hat{n}$).
For our purpose, which is to integrate the linearized Einstein equations,
it can be shown that only the following linear combination is relevant
\begin{equation}
\Psi \equiv \frac{ f_{\nu} \Psi_{\nu} + f_{\bar{\nu}} \Psi_{\bar{\nu}}}
{ f_{\nu} + f_{\bar{\nu}}}.
\end{equation}
Using the Boltzmann equations for $\Psi_{\nu}$ and $\Psi_{\bar{\nu}}$,
it is straightforward to show that the evolution of $\Psi$ (in Fourier
space and in the synchronous gauge, see \cite{MaBer}, eq. (40)) obeys
\begin{equation} \label{boltzmann}
\frac{\partial \Psi}{\partial \tau} +
i
\frac{Q}{\sqrt{Q^2 + a^2 M^2}}
(\vec{k}.\hat{n}) \Psi
+
\frac{d \ln ( f_{\nu} + f_{\bar{\nu}} )}{d \ln Q}
\left[
\dot{\eta} -
\frac{\dot{h}+6\dot{\eta}}{2}
(\hat{k}.\hat{n})^2 \right]=0~.
\end{equation}
This equation depends on $\xi$ only through the last term, which is
the gravitational source term.
In the case $\xi=0$, the quantity $(d \ln ( f_{\nu} )/d \ln Q)$ has a
simple interpretation: it is the $Q$-dependence of a planckian
perturbation of the phase space distribution. In other words, a shift
of the blackbody temperature $\Delta T / T(\vec{x}, \hat{n}, \tau)$
corresponds to a perturbation
\begin{equation}
\Psi (\vec{x}, Q, \hat{n}, \tau)
= - \frac{\Delta T}{T}(\vec{x}, \hat{n}, \tau)
\frac{d \ln ( f_{\nu} )}{d \ln Q}.
\end{equation}
Since the gravitational source term in the Boltzmann equation is
proportional to this quantity, the planckian shape is unaltered for
massless neutrinos, and also for massive neutrinos when they are still
relativistic (indeed, when $Q^2 \gg a^2 M^2$, the $Q$-dependence of
the Boltzmann equation (\ref{boltzmann}) vanishes in the second term,
and remains only in the third term). When $\xi\neq0$, the source term
in eq. (\ref{boltzmann}) is proportional to
\begin{equation} \label{dlnfdlnq}
\frac
{d \ln ( f_{\nu} + f_{\bar{\nu}} )}
{d \ln Q} =
-
\frac
{Q (1 + \mathrm{ch} \xi~\mathrm{ch} Q)}
{(\mathrm{ch} \xi + e^{-Q})(\mathrm{ch} \xi + \mathrm{ch} Q)}.
\end{equation}
When neutrinos are still relativistic, $\Psi$ is proportional to
this quantity, even if it cannot be simply interpreted in
terms of blackbody temperature perturbations.
We can now specify all the changes required in {\tt cmbfast}, first in
the case of massive degenerate neutrinos. As usual, $\Psi$ can be
expanded in a Legendre series: $\Psi = \sum_{l=0}^{\infty} (-i)^l (2l
+1) \Psi_l P_l$. It is easy to show that for each multipole $\Psi_l$,
the evolution equation and the initial condition are both identical to
those of the non-degenerate case, provided that we replace $(d \ln (
f_{\nu} )/d \ln Q)$ by eq. (\ref{dlnfdlnq}). So, in summary, one only
needs to modify the homogeneous phase-space distribution, its
logarithmic derivative with respect to $Q$, and the initial
calculation of $\Omega_{\nu}$. Also, in order to obtain a good
precision in the CMB anisotropy spectra, one must set $l=5$ for the
number of multipoles $\Psi_l$ to be time-integrated. For transfer
functions, the value $l=25$ proposed by the code is sufficient.
In the case of massless degenerate neutrinos, the situation is even
simpler. The $Q$-dependence of the Boltzmann equation can be
integrated away, just like in the non-degenerate case. For this
purpose, we must introduce the $Q$-independent variable $F_{\nu}$
\begin{equation}
F_{\nu} (\vec{k}, \hat{n}, \tau) \equiv
\frac{\int Q^3 dQ (f_{\nu} + f_{\bar{\nu}}) \Psi}
{\int Q^3 dQ (f_{\nu} + f_{\bar{\nu}})}
\equiv \sum_{l=0}^{\infty} (-i)^l (2l+1) F_{\nu l} P_l,
\end{equation}
and integrate by part the last term in equation (\ref{boltzmann}).
The multipoles $F_{\nu l}$ are exactly identical for degenerate and
non-degenerate massless neutrinos, because they share the same
evolution equations and initial conditions. So, the effect of $\xi$
arises only through the background quantities in \eq{background}
and is completely described by introducing an effective number of massless
neutrinos.
\section{Results}
\label{results}
First, as a consistency check, we compute CMB anisotropies and
transfer functions for different values of $\xi$, choosing a very
small mass $m_{\nu} \leq 0.001$ eV. We check that the results match
exactly those obtained with the unmodified version of {\tt cmbfast},
when the appropriate effective neutrino number $N_{eff}$ is specified.
The effect of $\xi$ and $m_{\nu}$ on the CMB anisotropy spectrum can
be seen in figure \ref{fig.CMB}. We choose a set of cosmological
parameters ($h=0.65$, $\Omega_b=0.05$, $\Omega_{\Lambda}=0.70$,
$\Omega_{CDM}=1-\Omega_b-\Omega_{\nu}-\Omega_{\Lambda}$,
$Q_{rms-ps}=18~\mu$K, flat primordial spectrum, no reionization, no
tensor contribution), and we vary $\xi$ from 0 to 5, both in the case
of massless degenerate neutrinos (solid lines) and degenerate
neutrinos with $m_{\nu}=0.07$ eV (dashed lines). Let us first comment
the massless case. The main effect of $\xi$ is to boost the amplitude
of the first peak\footnote{In fact, this is not true for very large
values of $\xi$. In such cases, recombination can take place still at
the end of radiation domination, and anisotropies are suppressed. For
our choice of cosmological parameters, this happens for $\xi \raise0.3ex\hbox{$\;>$\kern-0.75em\raise-1.1ex\hbox{$\sim\;$}} 7$,
but in such a case the location of the first peak is $l \raise0.3ex\hbox{$\;>$\kern-0.75em\raise-1.1ex\hbox{$\sim\;$}} 450$, and
the matter power spectrum is strongly suppressed.}. Indeed,
increasing the energy density of radiation delays matter-radiation
equality, which is known to boost the acoustic peaks \cite{Hu} (the
same explanation holds for the effect of $\Omega_{\Lambda}$ in flat
models). For the same reason, all peaks are shifted to higher
multipoles, by a factor $( (1 + a_{eq}/a_*)^{1/2} -
(a_{eq}/a_*)^{1/2})^{-1}$ \cite{Hu} ($a_{eq}$ is the scale factor at
equality, and increases with $\xi$, while the recombination scale
factor $a_*$ is almost independent of the radiation energy density).
Secondary peaks are then more affected by diffusion damping at large
$l$, and their amplitude can decrease with $\xi$.
In the case of degenerate neutrinos with $m_{\nu}=0.07$ eV, the
results are quite similar in first approximation. Indeed, the effects
described previously depend on the energy density of neutrinos at
equality. At that time, they are still relativistic, and identical to
massless neutrinos with equal degeneracy parameter. However, with a
large degeneracy, $\Omega_{\nu}$ today becomes significant: for
$\xi=5$, one has $\Omega_{\nu}=0.028$, i.e. the same order of
magnitude as $\Omega_b$. Since we are studying flat models,
$\Omega_{\nu}=0.028$ must be compensated by less baryons, cold dark
matter (CDM) or $\Omega_{\Lambda}$. In our example, $\Omega_b$ and
$\Omega_{\Lambda}$ are fixed, while $\Omega_{CDM}$ slightly
decreases. This explains the small enhancement of the first peak
compared to the massless case. Even if this effect is indirect, it is
nevertheless detectable in principle (even if one does not impose the
flatness condition, the effect of $\Omega_{\nu}$ will be visible
through a modification of the curvature). In figure \ref{fig.CMB}, for
$\xi=0$, the first peak maximum is enhanced by only 0.37\%, while for
$\xi=5$, there is an increase of 3.4\%, detectable by the future
satellite missions {\it MAP} and {\it Planck}, unless there are large
parameter degeneracies. It is well-known that such degeneracies are
generally removed when CMB and LSS data are combined for parameter
extraction \cite{Tegmark}.
We plot in figure \ref{fig.PK} the power spectrum $P(k)$ for the same
models as in figure \ref{fig.CMB}, normalized on large scales to
COBE. The effect of both parameters $\xi$ and $m_{\nu}$ is now to
suppress the power on small scales. Indeed, increasing $\xi$
postpones matter-radiation equality, allowing less growth for
fluctuations crossing the Hubble radius during radiation
domination. Adding a small mass affects the recent evolution of
fluctuations, and has now a direct effect: when the degenerate
neutrinos become non-relativistic, their free-streaming suppresses
growth of fluctuations for scales within the Hubble radius. For
non-degenerate neutrinos, this effect is known to reduce power on
those scales by a relative amount $\Delta P/P \sim 8 \Omega_{\nu} /
\Omega_0$ \cite{Huetal} (we introduced $\Omega_0 = 1 -
\Omega_{\Lambda}$). So, even with $m_{\nu}=0.07$ eV and $\xi=0$, it is
significant, especially at low $\Omega_0$. In the models of figure
\ref{fig.PK}, $P(k)$ decreases by $\sim 5$ \%, in agreement with the
theoretical prediction ($\Omega_{\nu}=1.8 \times 10^{-3}$,
$\Omega_0=0.3$). However, at $\xi=5$ (i.e. $\Omega_{\nu}=0.028$),
this effect is even larger: $P(k)$ decreases by a factor 2.2, instead
of an expected 1.7. This effect is likely related to the phase-space
distribution of neutrinos with a chemical potential: their average
momentum is shifted to larger values, making the free-streaming
suppression mechanism even more efficient.
Let us compare our results with those of previous works. The effect of
$\xi$ on the CMB for massless neutrinos and $\Omega_{\Lambda}=0$ is
the same as that one found in \cite{Sarkar}. We also agree with the
revised results in \cite{Kinney}.
\section{Comparison with observations}
\label{comparison}
Since the degeneracy increases dramatically the amplitude of the first
CMB peak, we expect large $\xi$ values to be disfavored in the case of
cosmological models known to predict a fairly high peak. On the other
hand, a high $\xi$ is likely to be allowed (or even favored) for
models that predict systematically a low peak, unless a large scalar
spectral index $n \geq 1.2$ ({\it blue tilt}) is invoked. For
instance, the degeneracy is likely to be favored by: (i) a large
contribution of tensor perturbations; (ii) a significant effect from
reionization; (iii) a low baryon density; (iv) a large $h$ ($h \geq
0.7$); (v) flat models with $\Omega_{\Lambda} \leq 0.6$; etc. For
such models, the peak amplitude can be boosted by $\xi$, keeping $n$
close to one, which is more natural from the point of view of
inflation. However, a careful case-by-case analysis is required, since
the effects of $\xi$ and $n$ on CMB and LSS spectra are far from being
equivalent. Our goal here is not to explore systematically all
possibilities, but to briefly illustrate how $\xi$ can be constrained
by current observations for flat models with different values of
$\Omega_{\Lambda}$. Recent results from supernovae
\cite{Perlmutter}, combined with CMB constraints, favor flat models
with $\Omega_{\Lambda} \sim 0.6-0.7$ \cite{Lineweaver}.
We choose a flat model with $h=0.65$, $\Omega_b=0.05$,
$Q_{rms-ps}=18~\mu$K, no reionization and no tensor contribution, and
look for the allowed region in the space of free parameters
($\Omega_{\Lambda},\xi,n$). The allowed region will not be defined
using a maximum likelihood analysis, but with the more conservative
technique called ``concordance'' by Wang et al. \cite{Wang}, which
consists in taking the intersection of regions allowed by each
experiment.
For simplicity, we take into account only a few constraints on the
matter power spectrum, known to be representative of the large amount
of available data: the value of $\sigma_8$ (the variance of mass
fluctuations in a sphere of radius $R=8 h^{-1} $Mpc) given for flat
models in \cite{ViaLid}, at 95 \% confidence level (CL)\footnote{The
Viana \& Liddle result \cite{ViaLid} is in very good agreement with an
independent derivation by Girardi et al. \cite{Borgani}.}; a
$\chi^2$ comparison with the STROMLO-APM redshift survey
\cite{Loveday}, at scales well within the linear regime, also with 95
\% CL\footnote{This confidence level stands for the goodness-of-fit of
the model: when $\chi^2$ is greater than some value, the probability
that we find the observed dataset, assuming the model to be valid, is
smaller than 5 \%. For the APM data, we have $9-3$ degrees of freedom,
and the limiting value is found in numerical tables to be
$\chi^2=12.5$.}; and finally, the constraint on bulk velocity at $R=
50 h^{-1}$ Mpc \cite{KolDek}, taking into account the cosmic variance.
Except for the updated $\sigma_8$ constraint, we use exactly the same
experimental tests as in \cite{LPS}, and refer the reader to this
paper for details. For CMB data, we perform a $\chi^2$ analysis based
on 19 experimental points and window functions, taking into account
the Saskatoon calibration uncertainty, in the way suggested by
\cite{LineBar}. The list of data that we use is given in \cite{LPS},
and again allowed regions correspond to 95 \% CL\footnote{Here we have
$19-3$ d.o.f.; the 95 \% CL is given by $\chi^2 = 26$.}. We do not
take into account the most recent experiments, for which window
functions are still unpublished; they are anyway in good agreement
with the data considered here.
We plot in figure \ref{fig.WIN} the LSS and CMB allowed regions in
($\xi$, $n$) parameter space, corresponding to
$\Omega_{\Lambda}=0,0.5,0.6,0.7$. For $\Omega_{\Lambda}=0.5-0.7$, the
LSS window just comes out of $\sigma_8$ limits. For
$\Omega_{\Lambda}=0$, the lower LSS constraint is from $\sigma_8$, and
the upper one from APM data. In the case of degenerate neutrinos with
$m_{\nu} = 0.07$ eV, the windows are slightly shifted at large $\xi$,
since, as we saw, the effect of $\xi$ is enhanced (dotted lines on the
figure). The CMB allowed regions do not show this distinction,
given the smallness of the effect and the imprecision of the data.
One can immediately see that LSS and CMB constraints on $n$ are shifted in
opposite direction with $\xi$: indeed, the effects of $\xi$ and $n$
both produce a higher CMB peak, while to a certain extent they
compensate each other in $P(k)$. So, for $\Omega_{\Lambda}=0.7$, a
case in which a power spectrum normalized to both COBE and $\sigma_8$
yields a too high peak\footnote{At least, for the values of the other
cosmological parameters considered here. This situation can be easily
improved, for instance, with $h=0.7$.}, a neutrino degeneracy can only
make things worst. On the other hand, for $\Omega_{\Lambda}=0.5-0.6$,
a good agreement is found up to $\xi \simeq 3$.
Let us finally consider the $\Omega_{\Lambda}=0$ case in which, after
COBE normalization of the power spectrum, there is a well known
discrepancy between the amplitude required by $\sigma_8$ and the shape
probed by redshift surveys: these two constraints favor different
values of $n$. We find that the neutrino degeneracy can solve this
problem with $\xi \raise0.3ex\hbox{$\;>$\kern-0.75em\raise-1.1ex\hbox{$\sim\;$}} 3.5$; however, the allowed window is cut at
$\xi \simeq 6$ by CMB data, and we are left with an interesting region
in which $\Omega_0=1$ models are viable. This result is consistent
with \cite{Sarkar}. However, current evidences for a low $\Omega_0$
Universe \cite{Perlmutter,Bahcall} are independent of the constraints
used here, so there are not many motivations at the moment to consider
this window seriously.
\section{Conclusions}
We have considered some cosmological implications of a large relic
neutrino degeneracy. We have shown that this degeneracy enhances the
contribution of massive neutrinos to the present energy density of the
Universe. For instance, neutrinos with a small mass $m_{\nu} \sim
10^{-2}$ eV can contribute significantly to $\Omega_0$, provided that
there is a large neutrino-antineutrino asymmetry.
Our main result is the computation of the power spectra of CMB
anisotropies and matter density in presence of a neutrino
degeneracy. We found, in agreement with \cite{Sarkar}, that it boosts
the amplitude of the first CMB peak, shifts the peaks to larger
multipoles, and supresses small scale matter fluctuations. These
effects follow the increase of neutrino energy density, that delays
matter-radiation equality.
We extended the calculation to the case of massive degenerate
neutrinos, and showed the results for a mass of $0.07$ eV, as
suggested by the Super-Kamiokande experiment. This mass has a small
effect on CMB anisotropies. Indeed, such light neutrinos are still
relativistic at recombination, but in presence of a degeneracy, they
can account for a substantial part of the density today, of order
$\Omega_{\nu} \sim 10^{-2}$. Also, we showed that small scale matter
fluctuations are much more suppressed when the degenerate neutrinos
are massive, because free-streaming of non-relativistic neutrinos is
more efficient when their average momentum is boosted by the chemical
potential.
We compared our results with observations, in the restricted case of a
flat universe with arbitrary ($\Omega_{\Lambda}, \xi, n$) and fixed
values of other cosmological parameters. We found that for
$\Omega_{\Lambda} \simeq 0.5 - 0.6$, a large degeneracy is allowed, up
to $\xi \simeq 3$. However, this upper bound is smaller than the value
$\xi \simeq 4.6$ needed to explain the generation of ultra-high energy
cosmic rays by the annihilation of high-energetic neutrinos on relic
neutrinos with mass $m_{\nu}= 0.07$ eV \cite{Gelmini}. We also tried
smaller values of $\Omega_{\Lambda}$, even if they are not favored by
combined CMB and supernovae data. It turns out that a large degeneracy
can account for both CMB and LSS constraints even with $\Omega_0 =1$,
provided that $3.5 \leq \xi \leq 6$. This analysis could be extended
to other cosmological models. For instance, the degeneracy is likely
to be compatible with a large contribution of tensor perturbations to
large scale CMB anisotropies.
Finally, it turns out that the degeneracy parameter and the mass of
degenerate neutrinos have effects within the level of detectability of
future CMB observations and redshift surveys, even with $m_{\nu} \sim
0.07$ eV. However, a careful analysis should be performed in order to
detect possible parameter degeneracy between $\xi$, $m_{\nu}$ and
other cosmological parameters.
\section*{Acknowledgements}
We thank S.~Borgani, A.~Masiero and A.Yu.~Smirnov for useful
discussions, as well as W.~Kinney and A.~Riotto for correspondence.
This work has been supported by INFN and by the TMR network grant
ERBFMRXCT960090.
|
\section{The ``single particle" technique}
The simplest way to observe the high energy component of GRBs
with a ground based experiment is using an air shower array
working in ``single particle" mode. Air shower arrays,
made of several counters spread over large areas, usually detect
air showers generated by cosmic rays of energy $E >$ 10-100 TeV;
the arrival direction of the primary particle is measured by
comparing the arrival time of the shower front
in different counters, and the
primary energy is evaluated by the number of secondary particles
detected.
Air shower arrays can be used in the
energy region E~$<$~1~TeV working in single particle mode,
i.e. counting all the particles hitting the individual detectors, independently
on whether they belong to a large shower or they are the lonely
survivors of small showers. Because of the cosmic ray spectrum steepness,
most of the events detected in this mode of operation
are in fact due to solitary particles from small
showers generated by 10-100 GeV cosmic rays.
Working in single particle mode an air shower array could in principle
detect a GRB in the energy range 1-10$^3$~GeV,
if the secondary particles generated by
the gamma-rays give a statistical significant excess of events
over the all-sky background due to cosmic rays.
The beauty of this technique consists in its extreme simplicity:
it is sufficient to count all the particles hitting the detectors
during fixed time intervals (i.e. every
second, or more frequently depending on the desired time resolution)
and to study the counting rate behaviour versus time,
searching for significant increases; the observation of an excess
in coincidence with a GRB seen by satellites would be an
unambiguous signature of the nature of the signal.
The single particle technique does not allow the
measurement of the energy and the direction of the primary gamma-rays,
because the number of particles (often only one per shower) is too small
to reconstruct the shower parameters. However it can allow
the study of the temporal behaviour of the high energy emission,
and, with some assumptions on the
spectral slope (possibly supported by satellite measurement at lower energy)
it can give an evaluation of the total power emitted at high energy.
Fig.1 shows the mean number $n_e$ of electrons and positrons
reaching the ground at different altitudes $h$ above
the sea level generated by a gamma-ray, as a function of the gamma-ray
energy, according to the Greisen analytical expression \cite{8}.
The curves are given for two different zenith angles of the primary
gamma-ray, $\theta$ = 0$^{\circ}$ and $\theta$ = 30$^{\circ}$.
The number of particles
strongly increases with the altitude, in particular at low energies,
and decreases at high zenith angles, in particular at low altitude.
Actually, due to the gamma-rays photoproduction,
a small number of muons is produced as well;
however, in the energy range $E=1-10^3$ GeV,
the number of muons is negligible compared with the
number of electrons and positrons (except for $h<$1~km, where they
are comparable) and they will not be considered in the following
discussion.
A possible way to increase the number of detectable particles is
to exploit the air shower photons, that are much more numerous than $e^{\pm}$.
Fig.2 shows the ratio $r_s=n_{ph}/n_e$
of the mean number of photons and the mean number of $e^{\pm}$,
as a function of the altitude.
The values have been obtained simulating electromagnetic cascades
induced by gamma-rays with a zenith angle $\theta$ = 30$^{\circ}$.
The detector is assumed with an energy threshold $E_{th}$=3 MeV.
The ratio $r_s$ range from $\sim$6 to $\sim$11 depending on the altitude
and on the primary gamma-ray energy. The conversion
of a fraction of photons in charged particles (by pair production
or Compton effect) would increase the detectable signal.
Placing a suitable layer of lead over the detector, a small fraction of
$e^{\pm}$ would be absorbed,
but a large number of photons would interact and
increase the GRB signal, by a factor depending on the converter
geometrical characteristics.
An interesting possibility to
avoid the absorption of $e^{\pm}$ is to adopt a thick scintillator
detector: it would work both as a target for photons and as a detector
for charged particles \cite{9}.
\subsection{The background}
The background is due to secondary charged particles
from cosmic rays air showers. The rate depends on the altitude $h$
and, to a lesser extent, on the geomagnetic latitude $\lambda$ of the observer.
The latitude effect
is almost negligible at sea level and increases with the altitude:
at $h$ = 5 km a.s.l. the
background at $\lambda$ $>$ 45$^{\circ}$ is about 1.6 times larger
then near the magnetic equator \cite{10}.
The dependence of the background on the altitude is more important.
Fig.3 shows the flux of $e^{\pm}$, $\mu^{\pm}$ and charged hadrons
as a function of the altitude above the sea level,
for an observer located near the geomagnetic equator,
obtained simulating with the CORSIKA code
the atmospheric cascades generated by protons and Helium nuclei
with an energy spectrum according to \cite{11}
and using a dipole model for the geomagnetic field.
The detector energy threshold is $E_{th}$=3 MeV.
The total background rate ranges from $\sim$250 to 2000 events m$^{-2}$ s$^{-1}$
increasing the altitude from 0 to 5000 m.
The rates obtained by the simulations are in good agreement with the
counting rates measured at 2000 m \cite{12} and 4300 m \cite{13}.
Most of the background is due to muons for 0 $< h <$ 3.5 km and to
$e^{\pm}$ at higher altitude
(note that a detector able to separate the electron and the muon
components would allow the reduction of the background by
the rejection of muons).
It's worth noting that the number of $e^{\pm}$ generated by a primary
gamma-ray (see Fig.1) increases much more rapidly with the altitude than the
background does.
This is due to two factors: {\it 1)} given a primary energy, the
electromagnetic shower generated by a proton penetrates more deeply in the
atmosphere than the gamma-ray shower does, {\it 2)}
the $\mu^{\pm}$ flux, that strongly
contributes to the background rate, is less dependent on the altitude
than the $e^{\pm}$ flux.
The possible conversion of secondary photons in detectable
particles (introduced in the previous section as a method
to increase the GRB signal) would produce an increase in the background as well.
Fig.3 also shows the number of photons reaching the ground
due to cosmic rays.
However the ratio $r_b$ of the number of photons and the
number of charged particles is
considerably smaller than the corresponding value $r_s$ for primary
gamma-rays ($r_b <3$ at any altitude, while $r_s \sim$ 6-11, as shown in Fig.2).
The background rate is not strictly constant and several mechanisms
are responsible for variations on different time scales
with amplitudes up to a few percent.
First of all, variations in the atmospheric pressure
affect the secondary particle flux:
an increase (decrease) of the ground level atmospheric pressure
results in a reduction (enhancement) of the background rate,
because of the larger (smaller) absorption of the electromagnetic
component. The pressure and the background rate are linearly
anti-correlated (e.g. the correlation coefficient
between the counting rate percent increase and the pressure variation
measured at 2000 m a.s.l. is $\sim -0.5\%$ per millibar \cite{14}).
By monitoring the pressure at the detector position it is possible
to correct the data for this effect.
The 24 hours anisotropy (due to the rotation of the Earth in the
interplanetary magnetic field), and the solar activity (e.g. large solar flares)
modulate the low energy primary cosmic ray flux.
Although the amplitudes of these variations are not negligible
(up to a few percent of the counting rate in some cases)
the time scales of these phenomena (hours) are too long to simulate
short duration events as GRBs.
Shorter variations in the single particle counting rate have been measured
in coincidence with strong thunderstorms and have been ascribed to the
effects of atmospheric electric fields on the secondary particles flux
\cite{14,15}; however the occurrences of these events are very rare
and also in this case the observed time scales
($\sim$ 10-15 minutes) are longer than the typical GRB duration.
From the experimental point of view,
more troublesome are possible instrumental effects, such as electrical noises,
that could simulate a GRB signal producing spurious increases
in the background rate.
Working in single particle mode require very
stable detectors, and a very careful and continuous monitoring of the
experimental conditions.
By comparing the counting rate of the single detectors
(i.e. the scintillator counters, in the case of an air shower array)
and requiring simultaneous and consistent variations of the rate
in all of them, it is possible
to identify and reject the noise events due to instrumental effects.
\section{Sensitivity to detect high energy GRBs}
The aim of this section is to quantify the
actual possibilities of observing
high energy GRBs using the single particle technique.
First, we present a simple model of high energy emission,
in order to define a GRB using a limitate set of parameters.
Subsequently, the response of the detector is evaluated
as a function of the burst parameters and detector features.
We conclude with an estimation of the possible rate of positive observations.
\subsection{Parametrization of the high energy emission}
In this simple model a GRB is assumed to release a
total amount of energy $L$
in photons of energy E $>$ 1 GeV;
the emission spectrum is assumed constant during the
emission time $\Delta t_e$ and represented by
$dN/dE \ = \ K \ E^{-\alpha }$ (number of photons emitted
per unit energy) up to a cutoff energy $E_{max}$.
The value of $E_{max}$ depends on the emission processes and the
physical conditions at the source;
a sharp cutoff at $E_{max}$ is probably unrealistic,
but for our purposes such a simple model is sufficient.
The total energy $L$ and the spectrum are related by
$L=K\int_{1~GeV}^{E_{max}}E^{-\alpha+1}dE$.
If the GRB source is located at a cosmological distance corresponding
to a redshift $z$, the observed burst duration is
$\Delta t=\Delta t_e (z+1)$ due to time dilation,
and the shape of the observed spectrum is determined by two factors:
\noindent
$a)$ the photons energies
are redshifted by a factor ($z+1$) due to the expansion of the universe
(the spectrum slope $\alpha$ doesn't change);
\noindent
$b)$ the high energy gamma-rays are absorbed in the intergalactic space through
$\gamma + \gamma \rightarrow e^+ e^-$ pair production
with starligth photons;
the probability of reaching the Earth for a photon emitted by a source
with a redshift $z$ is $e^{-\tau(E,z)}$, where
$\tau$ increases with $E$ and $z$.
The absorption becomes important
when $\tau >1$; according to \cite{16}, $\tau$ reaches the
value of unity at $z\sim$0.5 for $E$=100 GeV, and at $z\sim$2 for $E$=20 GeV.
Fig. 4 shows a spectrum with a slope
$\alpha=2$ affected by the absorption, for different $z$,
obtained according to \cite{16} up to 500 GeV
(at higher energy the curves are extrapolated).
As a consequence, the flux of photons at Earth is a power law
spectrum with the same slope of the emission spectrum
up to a certain energy depending on the distance, followed by a gradual
steepening, with a sharp end at the energy $E'_{max}=E_{max}/(z+1)$.
Assuming an isotropic emission,
the number of photons per unit area and unit energy
at the top of the atmosphere is
\begin{equation}
\frac{d\Phi_{\gamma}}{dE} = \frac{K}{4 \pi r^2} [E(z+1)]^{-\alpha} (z+1) e^{-\tau(E,z)}
\end{equation}
$r$ is the cosmological comoving coordinate \cite{17}
\begin{equation}
r=c\frac{zq_0+(q_0-1)(-1+\sqrt{2q_0z+1})}{H_0 q_0^2 (1+z)}
\end{equation}
$H_0$ is the Hubble constant and $q_0$ the deceleration parameter
(we will use $H_0$=75 km s$^{-1}$ Mpc$^{-1}$ and $q_0$=0.5,
i.e. flat universe with $\Omega=1$).
In this simple parametrizazion L, $z$, $\alpha$ and $E_{max}$ determine
the energy fluence of gamma-rays above 1 GeV at Earth,
$F=\int_{1~GeV}^{E'_{max}}\frac{d\Phi_{\gamma}}{dE} E dE$.
The possible range of variability of these parameters can be reasonably
deduced by lower energy measurements:
\noindent
\underline{$z$}:
up to now, the distance of GRBs host galaxies has been measured
for 6 events and the observed redshifts are respectively
0.8, 1.0, 1.1, 1.6, 1.6, 3.4 \cite{18,19,20,21,22,23}
(a further one, GRB980425, probably associated with a Supernova
explosion at $z$=0.0085 seems a peculiar and non standard event \cite{24});
basing on these data, we will take $z$ in the interval 0-4;
\noindent
\underline{$\alpha$}:
most of the events observed by BATSE show power law spectra
at energies $E>$100~KeV-1~MeV, with exponent 1.5$<\alpha<$2.5 \cite{25};
since the few GRBs detected by EGRET at energies $E>$30 MeV
also show a similar behaviour \cite{5},
one can expect a higher energy component with the same spectral
slope observed in the MeV-GeV energy region;
\noindent
\underline{$L$}:
assuming an isotropic emission, the energies released by these 5 bursts
range between $\sim$5~10$^{51}$ and $\sim$2~10$^{54}$ ergs
in the BATSE energy range $E>$20~KeV;
it is not unreasonable to assume the energy $L$ emitted above
1~GeV being of the same
order of magnitude (recall that a spectrum with a exponent $\alpha$=2
means equal amount of energy release per decade of energy);
\noindent
\underline{$E_{max}$}:
the value of the maximum energy of gamma-rays
depends on unknown physical conditions and in fact is one of the
quantity that we hope to measure with this technique;
we will consider $E_{max}$ as a parameter ranging in the interval
30~GeV-1~TeV.
\subsection{The GRB signal}
In this section we evaluate the response of a detector to a GRB,
as a function of the burst parameters and the detector characteristics.
A detector can be simply defined by the area
$A_d$ and the altitude $h$ above the sea level.
Note that the detector sensitivity does not depend on its
geometrical features, as the area of the single counters or their
relative positions, but only on the total sensitive area
(e.g. for an air shower array $A_d$ is the sum of the single counters areas).
The latitude of the detector geographic location will not be
considered due to its small effect on the background rate.
Given a GRB, defined by the parameters L, $z$, $\Delta t_e$,
$\alpha$ and $E_{max}$ previously discussed,
and with an arrival direction corresponding to a zenith angle $\theta$,
the flux of secondary particles (number of particles per unit
area) reaching the altitude $h$ above the sea level
induced by the gamma-rays of energy $E>$1~GeV is:
\begin{equation}
\Phi_{e^{\pm}} \ = \int_{1 \ GeV}^{E'_{max}}
\frac{d\Phi_{\gamma}}{dE} n_e(E,\theta,h) \ dE
\end{equation}
where $n_e(E,\theta,h)$ is the mean number of particles reaching the
altitude $h$ generated by a gamma-ray of energy $E$ and
zenith angle $\theta$.
As an example Fig.5 shows the particles flux as a function of the altitude,
produced by a vertical GRB with $L=10^{53}$ erg, for different distances.
The spectrum is assumed with a slope $\alpha$=2 and $E_{max}$=30 GeV and 1
TeV. The flux strongly increases with the altitude; at 5000 m it
is about 3 orders of magnitude higher than at sea level.
For distances less than $z \sim$1, the flux increases with $E_{max}$,
while at higher distances, due to the absorption of the most energetic
gamma-rays, the flux is almost independent on $E_{max}$.
The ground level flux decreases at higher zenith angles,
in particular at lower altitude (see Fig.1);
as an example at $h$=4000 m the flux at $\theta$=30$^{\circ}$
is $\sim$3 times lower than the vertical flux, while at $\theta$=50$^{\circ}$
the flux is reduced by almost 2 orders of magnitude.
The number of particles giving a signal in the detector
\footnote{
This is true assuming that all the particles hitting a detector
give a signal; actually in a standard air shower array,
two or more particles of the same shower hitting the same counter
give only one signal, due to the limited time resolution;
however, working in the $<$ 1 TeV energy range $n_e$ is so small
that the fraction
of showers giving two or more particle in the same counter is negligible.}
is $N_s \ = A_d $cos$\theta \Phi_{e^{\pm}}$.
The signal must be compared to the noise
\mbox{$\sigma _b=\sqrt{A_d \ B(h) \ \Delta t}$},
given by the background fluctuations during the time
$\Delta t=\Delta t_e (z+1)$,
where $B(h)$ is background rate (number of events per unit area and unit time).
A GRB is detectable with a statistical significance
of $n$ standard deviations if $N_s/\sigma _b > n$.
The minimum value of $n$ necessary to consider an excess as a significant
signal depends on the search strategy. In the case of a search
in coincidence with satellites, since the frequency of GRBs occurring
in the field of view of a ground based detector is about one event every
few days, a level of 4 standard deviations is sufficient to give a high
reliability to the observation. In the following discussion we set $n=$ 4.
A detector of a given area $A_d$ and altitude $h$ is therefore sensitive to
GRBs with a ground level particle flux
\begin {equation}
\Phi_{e^{\pm}} > \frac{n}{cos \theta} \sqrt{\frac{B(h) \Delta t}{A_d}}
\end {equation}
As an example, in Fig.5 the minimum detectable ground level flux
as a function of the altitude
is shown for a detector of area $A_d$=10$^3$ m$^2$ and a GRB with a
time duration $\Delta t$=1 s and zenith angle $\theta$ = 0$^{\circ}$.
The minimum flux ranges from $\sim$2 to $\sim$6 particles m$^2$ for altitudes
from the sea level to 5000 m.
Conversely, if one aims to be sensitive to a certain ground level
flux $\Phi_{e^{\pm}}$, the requirement for the minumum detector area
is
\begin {equation}
A_d > \frac{n^2 B(h) \Delta t}{\Phi_{e^{\pm}}^2 cos^2 \theta}
\end {equation}
To give an estimate of the magnitude of the largest
fluxes that one can reasonably expect,
we have extrapolated the power law fits of 15 GRBs spectra
obtained by the TASC instrument of EGRET \cite{5} up to a maximum energy
$E_{cut}$,
and we have calculated the secondary particle flux at the ground level
using equation (3) (setting $E'_{max}=E_{cut}$ and
$d\Phi_{\gamma}/dE$ equal to the EGRET spectra).
The maximum flux is given by GRB910709 ($\alpha$=1.76), which would
produce a flux $\Phi_{e^{\pm}} \sim$450(6500) particles m$^{-2}$ at
$h$=2000(5000) m
assuming $E_{cut}$=1 TeV, and $\Phi_{e^{\pm}} \sim$12(420) particles m$^{-2}$
for $E_{cut}$=30 GeV.
In total, the number of GRBs (out of 15) giving
a flux $\Phi_{e^{\pm}}>$ 10 particle m$^{-2}$ at the altitude $h$=2000(5000) m
is $N=$ 3(10) for $E_{cut}$=1 TeV, and $N=$1(4) for $E_{cut}$=30 GeV.
As expected the ground level signals are very sensitive to the maximum energy
$E_{cut}$, that is determined by the maximum emitted energy at the source
and by the absorption in the intergalactic space.
Our analysis of the EGRET events indicates that even if the GRBs spectra
at Earth do not extend to very high energy, the most powerful events
can give ground level fluxes at mountain altitudes
$\Phi_{e^{\pm}} >$ 10 particles m$^{-2}$.
The minimum sensitive area required to detect a flux of 10 particles m$^{-2}$
can be deduced by the expression (5)
using the background estimates given in Fig.3.
For example, at $h$=4000 m,
\mbox{$A_{min}=214 \frac{\Delta t}{1 s} \frac{1}{cos^2 \theta}$ m$^2$.}
So far we have expressed the sensitivity of the single particle
technique in terms of the ground level particles flux.
To discuss the sensitivity
in terms of the primary gamma-rays flux at the top of the atmosphere,
it is useful to introduce $F_{min}$,
the minimum energy fluence of gamma-rays above 1 GeV
required to give a detectable signal.
We considered a detector of area $A_d$=10$^3$ m$^2$
and a GRB with a zenith angle $\theta$ = 0$^{\circ}$,
$\Delta t$=1 s, and $z$ sufficiently small to make the distortions of the
spectrum (absorption and redshift) negligible.
Fig.6 shows the minimum energy fluence $F_{min}$
in the range 1 GeV-$E_{max}$
as a function of $h$, for different spectral parameters.
If the detector is located at very high mountain altitude (h$>$ 4000 m)
fluences of few 10$^{-5}$ erg cm$^{-2}$ are observable
also from soft spectra ($\alpha$=2.5, $E_{max}$=30 GeV),
while fluences of few 10$^{-6}$ erg cm$^{-2}$ are detectable only
from hard spectra ($\alpha$=1.5, $E_{max}$=1 TeV).
The minimum observable fluence for a detector with a generic
area $A_d$ = $k$ 10$^3$ m$^2$ and for a GRB with a generic time duration
$\Delta t=T$ s, scales as $\sqrt{T/k}$.
It is worth noting that in the evaluation of the sensitivity
we have not taken into account the possibility
to increase the GRB signal by converting a fraction of photons in detectable
charged particles as discussed in Section 2.
\subsection{Expected rate of observations}
A rough evaluation of the expected rate of positive observations,
as a function of the unknown spectral parameters at high energies,
can be performed by making assumptions on the luminosity function and
the spatial distribution of GRBs in the universe.
As a first step we evaluate the probability to detect a GRB
with a given luminosity\footnote{
The luminosity is usually defined as the amount of energy emitted
per unit time, while here we call luminosity
the energy emitted during the total emission time $\Delta t_e$.}
L as a function of $z$.
Due to the absorption, the luminosity $L$ required to observe a GRB
increases with the
distance more rapidly than what is expected by simple geometrical effects.
Fig.7 shows the minimum luminosity $L$ necessary to make a burst
detectable
by a 10$^3$ m$^2$ detector located at different altitudes,
as a function of the redshift $z$.
The burst is assumed at a zenith angle $\theta$=0$^{\circ}$,
with a time duration
$\Delta t_e$=1 s, $\alpha$=2 and $E_{max}$=30 and 1000 GeV.
Obviously, for a detector of a generic area $A_d$ = $k$ 10$^{3}$ m$^2$
the minimum observable luminosity scales as 1/$\sqrt{k}$.
From these curves one can deduce the maximum distance
observable by the detector, for a given $L$.
As an example, if $L$ = $10^{51}$(10$^{54}$) erg and $E_{max}$=1 TeV,
a 10$^3$ m$^2$ detector at the
sea level could see the burst if the source is
located within a distance $z \sim$ 0.003(0.4)
while a detector at $h$=4 km could see up to $z \sim$ 0.2 (3.0).
As a matter of fact not all GRBs with a given $L$ are
observable if $z$ is less than a certain value: the detection depends
on the zenith angle $\theta$. A correct approach to the problem
is to calculate the probability $P(L,z)$ to observe a burst of
luminosity $L$ and distance $z$ taking into account all possible
arrival directions
\begin{equation}
P(L,z) = \frac{1}{4 \pi} \int_{0}^{2\pi} d\phi \int_{0}^{\pi} J(L,z,\theta) sin\theta d\theta
\end{equation}
where $J(L,z,\theta)$=1 if the burst is detectable, i.e. if the
statistical significance
of the signal is larger then the required value, otherwise
$J(L,z,\theta)$=0.
As a second step we evaluate
the fraction $\epsilon(L)$ of observable events over the total number of
GRBs of luminosity $L$ distributed in the universe.
For this purpose one should know
the form of the GRBs density $n(z,L)$ (number of bursts
per unit volume and unit time). We will adopt the simplest model:
$a)$ GRBs are distributed in the universe with constant
density and frequency
in a comoving coordinate frame up to a maximum distance $z_{max}$,
$b)$ GRBs show no evolution, i.e. the luminosity distribution is
independent on $z$.
With these assumptions
$n(z,L)\propto (1+z)^3/(1+z)$;
the term $(1+z)^3$ takes into account that the universe
at a time corresponding to a redshift $z$ was smaller than now by a factor
$(1+z)$; the term $1/(z+1)$ represents the frequency decrease due
to time dilation. The fraction of observable events up within the distance
$z_{max}$ is
\begin{equation}
\epsilon(L) = \frac{\int_{0}^{z_{max}} P(L,z) (1+z)^2 \frac{dV}{dz} dz}
{\int_{0}^{z_{max}} (1+z)^2 \frac{dV}{dz} dz}
\end{equation}
where
$dV(z)$ is the volume element, according to \cite{17}
\begin{equation}
dV(z) = \frac{4 \pi c^3}{(1+z)^6 H_0^3 q_0^4} (1+2q_0z)^{-1/2}
[zq_0+(q_0-1)(-1+\sqrt{2q_0z+1})]^2 dz
\end{equation}
Fig.8 shows $\epsilon(L)$ for a 10$^3$ m$^2$ detector at different
altitudes, calculated setting $z_{max}$=4.
The burst is assumed with an emission time
$\Delta t_e$=1 s, a slope $\alpha$=2 and $E_{max}$=30, 100 and 1000 TeV.
The corresponding fraction of observable events for a detector area
$A_d$ = $k$ 10$^{3}$ m$^2$ is $\epsilon(L \sqrt{k})$.
It is interesting to note that, for a given detector altitude,
a GRB with $E_{max}$=100 GeV becomes more easily observable than
a GRB with the same luminosity and $E_{max}$=1 TeV,
if $L$ is larger than a certain value; this is due to the fact
that increasing the luminosity one can observe at
larger distances and consequently the absorption of high energy photons
becomes more important.
The final step is the calculation of the total rate
of observable events, considering all the possible luminosities $L$
\begin{equation}
R_{obs} = R_{tot}\int_{L_{min}}^{L_{max}} \frac{dN(L)}{dL} \epsilon(L) dL
\end{equation}
where $R_{tot}$ is the total bursts rate in the universe within a
distance $z_{max}$ and dN(L)/dL is the luminosity distribution of GRBs.
A rough evaluation of $R_{tot}$ is given from BATSE data.
Since BATSE observes $\sim$1 burst per day, with a detection
efficiency of $\sim$0.3, the rate deduced from BATSE
obervations is $R_B\sim$1000 y$^{-1}$. In the assumption that BATSE
can observe GRBs up to a distance $z \sim$ 4, we set $R_{tot}=R_B$.
The luminosity distribution is a more troublesome topic:
in the KeV-MeV energy region
a large number of studies has been done in order to deduce a luminosity
function suitable to reproduce the flux distribution of BATSE
bursts but a general agreement about the subject doesn't exist.
Different forms of $N(L)$ could reproduce the BATSE data and
so far there exist too few luminosity measurements to put
stringent constraints on the distribution.
One point is clear: the 5 events whose distance has been measured
show that GRBs are not standard candles and the luminosity distribution
at low energy is a very broad function, ranging over almost 4 orders of
magnitude.
Lacking more compelling data we will assume the luminosity
function in the form $dN(L)/dL=C L^{-\beta}$
with $L_1 < L < L_2$; tentatively,
we set $L_1 = 10^{51}$ erg and $L_2 = 10^{55}$ erg,
and consider $\beta$ as parameter ranging in the interval 1-4.
Fig.9 shows the expected rate of observation as a function of $\beta$
for a 10$^3$ m$^2$ detector at different altitudes.
The burst spectrum is assumed with $\alpha$=2 and
$E_{max}$=30, 100 and 1000 GeV.
The rates are given for the emission times $\Delta t_e$=1 s and $\Delta t_e$=10 s.
The curves show that a 10$^3$ m$^2$ detector located at
$h \sim$ 4 km, could observe a number of bursts per year ranging from $\sim$ 1 to
20-30 if the slope of the luminosity function is not steeper than
2-2.5.
\section{Conclusions}
The single particle technique provides a simple method of observing
GRBs in the energy range 1-10$^3$ GeV using large air shower
arrays located at high mountain altitude.
The observation of a significant excess in the counting rate in coincidence
with a burst seen by a satellite
would provide, in a very simple and economical way,
important informations on the GRB high energy component.
Besides the study of the temporal behaviour of the high energy emission,
the single particle observation could allow the evaluation of
the total energy $L$ emitted above 1 GeV (or equivalently
the maximum energy of the spectrum $E_{max}$) provided that
the spectral slope is deduced by a complementary
lower energy observation and the redshift is measured with optical instruments.
The expected rate of observation depends on bursts parameters that are
so far poorly known, as the high energy spectrum shape,
the distribution of the sources in the universe, and in
particular the luminosity function.
However, making reasonable assumptions,
we conclude that a detector of $A_d \sim 10^3$ m$^2$ at
an altitude of $\sim$ 4000 m above the sea level
could be expected to detect
at least few events per year if
the GRB spectrum extends up to $\sim$ 30 GeV
and the luminosity function slope is not too large.
An attempt to detect high energy GRBs using the
single particle technique was performed
by G.Navarra about 15 years ago using a
small detector located at Plateau Rosa at 3500 m a.s.l.\cite{26}.
At present two experiments are using this technique:
the 350 m$^2$ air shower array EASTOP,
working at 2000 m a.s.l. at the Gran Sasso Laboratory (Italy)
and INCA, a 48 m$^2$ detector
located at Mount Chacaltaya (Bolivia) at 5200 m a.s.l.
EASTOP has been searching high energy
GRBs since 1991, showing the possibility
to perform the single particle measurement with the
necessary stability over long periods of time, monitoring
with very good reliability short time fluctuations in the counting rate.
In a search for high energy gamma-rays in coincidence
with about 300 BATSE events detected during 1992-1998, EASTOP
obtained upper limits to the 10 GeV-1 TeV fluence
as low as $F \sim 10^{-4} - 5 \ 10^{-3}$ erg cm$^{-2}$,
for small zenith angle GRBs ($\theta<30^{\circ}$)\cite{27}.
In a similar search performed with about 2 years of data,
INCA obtained upper limits to the 1 GeV-1 TeV fluence
$F \sim$ 5 $10^{-5}-10^{-4}$ erg cm$^{-2}$,
also for small zenith angle GRBs\cite{28,29}. The limited area of INCA is
highly compensated by its extreme altitude.
The single particle technique has been adopted as well by the CYGNUS
collaboration using water Cerenkov detectors, with a total sensitive
area of $\sim$210 m$^2$ at an altitude of 2130 m. They reported
upper limits on the 1~GeV-30~TeV energy fluence $F\sim
10^{-4} - 2 \ 10^{-1}$ erg cm$^{-2}$ s$^{-1}$,
in coincidence with 9 strong GRBs with zenith angles up to 60$^{\circ}$
\cite{30}.
Good possibilities of positive detections
come from the new generation air shower arrays
consisting of very large sensitive surfaces, as ARGO and MILAGRO,
conceived to detect small air showers with the aim of observing
gamma-ray sources at energy E $<$ 1 TeV.
ARGO, under construction at the
Yangbajing Laboratory (Tibet) at 4300 m, will consist of a carpet of 7500 m$^2$
of Resistive Plate Chambers, covered by a layer of lead 0.5 cm thick to
convert a fraction of shower photons and
increase the number of detectable particles \cite{31}.
MILAGRO, a Cerenkov detector using a 4800 m$^2$ pond of water,
is located at 2600 m, near Los Alamos. An interesting feature
of the water Cerenkov technique is the possiblity to perform a calorimetric
measurement of the shower particles and to detect most of
the photons, due to their interactions in the water \cite{30,32}.
The two future experiments have about the same sensitivity in observing
GRBs by using the single particle technique. The lower altitude
of MILAGRO, that reduces its sensitivity by a factor $\sim$3 with
respect to ARGO,
is compensated by the higher capability of detecting the shower photons.
If the assumptions that we have made on high energy GRBs are correct,
both the experiments likely will detect at least few events per year.
\vspace{0.4 cm}
{\bf\large Acknowledgements}
\vspace{0.4 cm}
We are grateful to P.Lipari, G.Navarra, O.Saavedra and P.Vallania
for helpful discussions and comments.
\newpage
|
\section{\@startsection {section}{1}{\z@}{-8.5ex plus -1ex minus
-.2ex}{3.3ex plus .2ex}{\large\bf\centering}}
\def\subsection{\@startsection{subsection}{2}{\z@}{-3.25ex plus
-1ex minus -.2ex}{1.5ex plus .2ex}{\bf}}
\def\subsubsection{\@startsection{subsubsection}{3}{\z@}{-3.25ex plus%
-1ex minus -.2ex}{1.5ex plus .2ex}{\sl}}
\renewcommand{\thefootnote}{\alph{footnote}}
\newcommand{\br_{\hs{-5pt} I}}{\biggr)_{\hspace{-5pt} I}}
\newcommand{\vs{1cm} \hs{5cm} \vdots \vs{1cm} }{\vspace{1cm} \hspace{5cm} \vdots \vspace{1cm} }
\newcommand{\vspace{1cm}\noindent{\bf\large{Acknowledgements}}}{\vspace{1cm}\noindent{\bf\large{Acknowledgements}}}
\newcommand{\eq}[1]{eq.(\ref{#1})}
\newcommand{{\boldsymbol \phi}}{{\boldsymbol \phi}}
\newcommand{{\boldsymbol \lambda}}{{\boldsymbol \lambda}}
\newcommand{{\boldsymbol \alpha}}{{\boldsymbol \alpha}}
\newcommand{\real}[1]{\text{Re}(#1)}
\newcommand{\imaginary}[1]{\text{Im}(#1)}
\newcommand{\text{sign}}{\text{sign}}
\newcommand{\mathbb{Z}}{\mathbb{Z}}
\newcommand{\bp}{C}
\newcommand{\e}{{\boldsymbol{e}}}
\newcommand{\er}{\boldsymbol{\epsilon}}
\newcommand{\kextra}{M}
\renewcommand{\topfraction}{1}
\renewcommand{\bottomfraction}{1}
\renewcommand{\textfraction}{0}
\theoremstyle{plain}
\newtheorem{theorem}{Theorem}[section]
\newtheorem{proposition}[theorem]{Proposition}
\newtheorem{lemma}[theorem]{Lemma}
\newtheorem{corollary}[theorem]{Corollary}
\theoremstyle{definition}
\newtheorem{definition}{Definition}[section]
\begin{document}
\begin{titlepage}
\vspace*{-2cm}
\begin{flushright}
KCL-MTH-99-10\\
DTP--99/23\\
hep-th/9904002 \\
\end{flushright}
\vspace{0.3cm}
\begin{center}
{\Large
{\bf Particle Reflection Amplitudes\\
\vspace{2mm}
in $a_n^{(1)}$ Toda Field Theories}}\\
\vspace{1cm}
{\large \bf G.\ W.\ Delius\footnote{\noindent E-mail:
{\tt <EMAIL>},~~~~home page:
{\tt http://www.mth.kcl.ac.uk/\~{}delius/}} and
G.\ M.\ Gandenberger}\footnote{\noindent E-mail:
{\tt <EMAIL>}}\\
\vspace{0.3cm}
{${}^a$\em Department of Mathematics\\
King's College London\\
Strand, London WC2R 2LS, U.K.}\\
\vspace{0.3cm}
{${}^b$\em Department of Mathematical Sciences\\
Durham University\\
Durham DH1 3LE, U.K.}\\
\vspace{1cm}
{\bf{ABSTRACT}}
\end{center}
\begin{quote}
We determine the exact quantum particle reflection amplitudes
for all known vacua of $a_n^{(1)}$ affine Toda theories
on the half-line with
integrable boundary conditions.
(Real non-singular vacuum solutions are known for about half
of all the classically integrable boundary conditions.)
To be able to do this we use the fact
that the particles can be identified with the
analytically continued breather solutions, and that the real vacuum
solutions are obtained by analytically continuing stationary
soliton solutions.
We thus obtain the particle reflection amplitudes
from the corresponding breather reflection amplitudes. These
in turn we calculate by bootstrapping from soliton reflection
matrices which we obtained as solutions of the boundary
Yang-Baxter equation (reflection equation).
We study the pole structure of the
particle reflection amplitudes and uncover an
unexpectedly rich spectrum of excited boundary states, created by
particles binding to the boundary. For $a_2^{(1)}$ and $a_4^{(1)}$
Toda theories we calculate the reflection amplitudes
for particle reflection off all these excited boundary states.
We are able to explain all physical strip poles in these
reflection factors either in terms of boundary bound states or a
generalisation of the Coleman-Thun mechanism.
\end{quote}
\vfill
\end{titlepage}
\section{Introduction and Overview}
The study of two-dimensional integrable quantum field theories is
interesting because it allows one to calculate exactly things like
the quantum behaviour of solitons,
the weak--strong coupling duality,
the effects of space-time boundaries, etc. Exact
studies in two dimensions allow one to speculate more confidently
about these phenomena also in higher dimensions.
A particularly rewarding class of integrable quantum field
theories to study are the affine Toda field theories (ATFTs). Some
of their nice features include: they
exhibit weak--strong coupling duality,
they have quantum group symmetries \cite{ber},
they have solitons and breathers \cite{hol,oli}, and,
most importantly for this paper,
they have a rich set of integrable boundary conditions
\cite{corri94,bowco95}.
For the particles of real coupling affine Toda theories without
boundaries it
has been possible to find the exact S-matrices, describing the
evolution of arbitrary asymptotic incoming particle states into
asymptotic outgoing particle states \cite{Arin,brade90,chri,deliu92}.
It has been a longstanding challenge to
extend these results to the half-line with integrable boundary
conditions.
We will only deal with $a_n^{(1)}$ Toda theory in this paper.
The classical equation of motion for the $n$-component bosonic
field of $a_n^{(1)}$ Toda theory is
\begin{equation}
\partial_t^2{{\boldsymbol \phi}}-\partial_x^2{{\boldsymbol \phi}}+\frac{m^2}{\beta}\sum_{i= 0}^n
{\boldsymbol \alpha}_ie^{\beta{\boldsymbol \alpha}_i\cdot {\boldsymbol \phi}}= 0.
\label{phieom}
\end{equation}
The $\alpha_i$, $i=1,\cdots,n$ are the simple roots of the Lie algebra
$a_n=\text{sl}_{n+1}$ and $\alpha_0$ is minus the highest root.
We will use units so that the mass scale $m= 1$.
Note that the coupling constant
$\beta$ could be removed from the
equations of motion by rescaling the field and
therefore the coupling
constant plays a role only in the quantum theory.
It has been discovered in \cite{corri94,bowco95} that
this equation of motion
can be restricted to the left half-line $x<0$ without
losing integrability if one imposes a boundary condition at
$x= 0$ of the form
\begin{equation}\label{bcond}
\left.\beta\,\partial_x{\boldsymbol \phi}+\sum_{i= 0}^n \bp_i\,{\boldsymbol \alpha}_i
e^{\beta{\boldsymbol \alpha}_i\cdot{\boldsymbol \phi}/2}\right|_{x= 0}= 0,
\end{equation}
where the boundary parameters $C_i$
satisfy either
\begin{equation}
C_i = 0\;, \hspace{1cm} (i=0,1,\dots,n)\;,
\end{equation}
which gives the Neumann boundary condition, or
\begin{equation}
C_i = \pm 1\;, \hspace{1cm} (i=0,1,\dots,n)\;, \label{boundcond}
\end{equation}
which will be denoted as the (++...-- --...) boundary
conditions.
The original ideas needed to determine the full S-matrix in the
presence of integrable boundary conditions are due to
Cherednik \cite{chere84},
Ghoshal and Zamolodchikov\cite{ghosh94} and Fring and
K\"{o}berle\cite{fring94}.
As a consequence of the higher-spin symmetries of affine Toda theory
all amplitudes factorise into
products of the two-particle scattering amplitudes and the
single-particle reflection amplitudes.
These amplitudes in turn are strongly restricted by the
requirements of real analyticity, unitarity, crossing symmetry and
the bootstrap equations. We recall some details regarding this in
section \ref{sect:boot}.
If one knows the spectrum of
the integrable theory under investigation, these constraints are
usually strong enough to determine the amplitudes uniquely. This
explains the success in finding the S-matrices for affine Toda theories on the
whole line, because one can read off the particle spectrum directly
from the Lagrangian. In a theory on the half-line however a large
number of boundary bound states can exist in addition to the
particle states.
Unfortunately there is
no obvious way how to read off the spectrum of boundary states
directly from the Lagrangian.
This is the reason why many previous attempts to determine the
reflection amplitudes relying only on the above mentioned constraints
have failed\footnote{See however \cite{corri94,corri94b} in
which conjectures
were based on calculations of boundary solutions.
Some of these conjectures we will confirm to be correct.}.
We will see below that the correct reflection factors
have a rather intricate boundary state spectrum.
The new approach to finding the particle reflection amplitudes
which we will use in this paper has
been proposed in \cite{gande98b}. It is based on an earlier discovery
that the lowest breather states in imaginary coupling
Toda theory can be identified with the fundamental particles in
the real coupling theory in the sense that the particle
S-matrices are obtained from the breather \mbox{S-matrices} by
analytic continuation in the coupling constant \cite{gande95,gande96}.
Previously this phenomenon was only known to be exhibited by the
sine-Gordon model.
In \cite{gande98b} it was
proposed that the same should also hold for the reflection
amplitudes.
Analytic continuation of the breather reflection amplitudes should
give the particle reflection amplitudes.
The breather reflection amplitudes are determined by
bootstrap from the reflection amplitudes of the constituent solitons.
The soliton reflection amplitudes in turn can be obtained by
solving the reflection equation which is the analogue
of the Yang-Baxter equation for reflection amplitudes.
In \cite{gande98b} the reflection equation for $a_2^{(1)}$
Toda theory was solved and the particle
reflection amplitudes for
$a_2^{(1)}$ Toda theory with Neumann or $(+++)$ boundary condition
were calculated.
The results are briefly recalled
and then generalised to $a_n^{(1)}$ Toda theory with arbitrary $n$
in section \ref{sect:an}.
We obtain the reflection amplitudes for the uniform $(++\cdots ++)$
boundary condition in \eq{Ka} and find that they
confirm an early conjecture made in \cite{corri94}.
The reflection amplitudes for the Neumann
boundary condition, which we conjecture to be obtained by duality,
are given in \eq{KNeum}.
A prerequisite for progress towards deriving the reflection
amplitudes
for the other integrable boundary conditions was an understanding
of their classical vacuum solutions.
Bowcock \cite{bowco96b} determined these vacuum solutions for
$a_n^{(1)}$ Toda theories.
This study of classical solutions was simplified and generalised
in \cite{deliu98}. It was found that for about half of all integrable
boundary conditions in $a_n^{(1)}$ Toda theory real-valued vacuum solutions
can be obtained by analytically continuing a solution describing
a stationary soliton in front of the
boundary. We therefore call these boundary conditions `solitonic'.
Some details of this are given in section \ref{sect:vacuum}.
With this knowledge it is then easy to obtain the particle reflection
amplitudes for the vacuum states for these boundary conditions.
This is done in section \ref{sect:mixed}. We find that the
boundary conditions fall into $\lfloor(n+1)/2\rfloor$ equivalence
classes (where $\lfloor k\rfloor$ is the integer part of $k$).
Namely all boundary conditions whose vacuum solutions are
obtained from solitons in the same multiplet or from solitons in
the conjugate multiplet give rise to the same reflection
amplitudes. The result for the reflection amplitudes can be found
in eqs. \eqref{Kmixed}, \eqref{statsolfactor}.
In section \ref{sect:poles} we study the pole structure of these
reflection amplitudes.
This constitutes most of the work in this paper. We discover that
the vacuum reflection factors have simple poles which imply the
existence of excited boundary states. We can obtain the amplitudes
for reflection off these excited boundaries by using the
boundary bootstrap equations.
These in turn have new poles implying the existence of yet more
excited boundary states.
We have performed this analysis in detail
for the cases of the $a_2^{(1)}$
and $a_4^{(1)}$ Toda theories. In each of these cases we
end up with an array of
boundary states. The number of excited boundary states grows as
the coupling constant becomes smaller. As an example we have
plotted in figures \ref{spectrum1} and \ref{spectrum2} the
spectrum of boundary states for the solitonic boundary conditions
of $a_2^{(1)}$ affine Toda theory at two different values of the
coupling constant
$B=\frac{\beta^2}{2\pi}/(1+\frac{\beta^2}{4\pi})$. In
figure~\ref{spectrum1} we have also indicated which particle
has to bind to the
boundary to go from one state to the other by labels on
the lines connecting the states. A more illuminating animated plot
of the spectrum can be found on this paper's web site at
{\tt http//www.mth.kcl.ac.uk/\~{}delius/pub/reflection.html}.
\begin{figure}
\begin{center}
\unitlength 0.8mm
\linethickness{0.4pt}
\begin{picture}(151.00,60.00)(0,50)
\put(80.00,100.00){\line(-1,-2){10.00}}
\put(70.00,80.00){\line(1,-2){10.00}}
\put(80.00,60.00){\line(1,2){10.00}}
\put(90.00,80.00){\line(-1,2){10.00}}
\put(100.00,65.00){\line(-2,3){10.00}}
\put(80.00,60.00){\line(4,1){20.00}}
\put(100.00,65.00){\line(4,3){20.00}}
\put(120.00,80.00){\line(1,1){20.00}}
\put(80.00,60.00){\line(-4,1){20.00}}
\put(60.00,65.00){\line(-4,3){20.00}}
\put(40.00,80.00){\line(-1,1){20.00}}
\put(60.00,65.00){\line(2,3){10.00}}
\put(80.00,100.00){\circle*{4.00}}
\put(90.00,80.00){\circle*{4.00}}
\put(70.00,80.00){\circle*{4.00}}
\put(20.00,100.00){\circle*{4.00}}
\put(40.00,80.00){\circle*{4.00}}
\put(60.00,65.00){\circle*{4.00}}
\put(80.00,60.00){\circle*{4.00}}
\put(100.00,65.00){\circle*{4.00}}
\put(120.00,80.00){\circle*{4.00}}
\put(140.00,100.00){\circle*{4.00}}
\put(11.00,61.00){\vector(0,1){49.00}}
\put(140.00,52.00){\line(0,-1){4.00}}
\put(130.00,52.00){\line(0,-1){4.00}}
\put(120.00,52.00){\line(0,-1){4.00}}
\put(110.00,52.00){\line(0,-1){4.00}}
\put(100.00,52.00){\line(0,-1){4.00}}
\put(90.00,52.00){\line(0,-1){4.00}}
\put(80.00,52.00){\line(0,-1){4.00}}
\put(70.00,52.00){\line(0,-1){4.00}}
\put(60.00,52.00){\line(0,-1){4.00}}
\put(50.00,52.00){\line(0,-1){4.00}}
\put(40.00,52.00){\line(0,-1){4.00}}
\put(30.00,52.00){\line(0,-1){4.00}}
\put(20.00,52.00){\line(0,-1){4.00}}
\put(80.00,104.00){\makebox(0,0)[cb]{\shortstack{\scriptsize{$b_{1,1}$}}}}
\put(140.00,104.00){\makebox(0,0)[cb]{\shortstack{\scriptsize{$b_{-3,3}$}}}}
\put(20.00,104.00){\makebox(0,0)[cb]{\shortstack{\scriptsize{$b_{3,-3}$}}}}
\put(36.00,77.00){\makebox(0,0)[rt]{\shortstack{\scriptsize{$b_{-2,2}$}}}}
\put(124.00,77.00){\makebox(0,0)[lt]{\shortstack{\scriptsize{$b_{2,-2}$}}}}
\put(65.00,80.00){\makebox(0,0)[rc]{\shortstack{\scriptsize{$b_{0,1}$}}}}
\put(95.00,80.00){\makebox(0,0)[lc]{\shortstack{\scriptsize{$b_{1,0}$}}}}
\put(60.00,60.00){\makebox(0,0)[ct]{\shortstack{\scriptsize{$b_{-1,1}$}}}}
\put(100.00,60.00){\makebox(0,0)[ct]{\shortstack{\scriptsize{$b_{1,-1}$}}}}
\put(80.00,55.00){\makebox(0,0)[ct]{\shortstack{\scriptsize{$b_{0,0}$}}}}
\put(88.00,91.00){\makebox(0,0)[lc]{\shortstack{\scriptsize{$2$}}}}
\put(77.00,71.00){\makebox(0,0)[lc]{\shortstack{\scriptsize{$2$}}}}
\put(98.00,72.00){\makebox(0,0)[lc]{\shortstack{\scriptsize{$2$}}}}
\put(72.00,91.00){\makebox(0,0)[rc]{\shortstack{\scriptsize{$1$}}}}
\put(62.00,72.00){\makebox(0,0)[rc]{\shortstack{\scriptsize{$1$}}}}
\put(83.00,71.00){\makebox(0,0)[rc]{\shortstack{\scriptsize{$1$}}}}
\put(68.00,60.00){\makebox(0,0)[ct]{\shortstack{\scriptsize{$1$}}}}
\put(46.00,70.00){\makebox(0,0)[ct]{\shortstack{\scriptsize{$1$}}}}
\put(27.00,89.00){\makebox(0,0)[rt]{\shortstack{\scriptsize{$1$}}}}
\put(90.00,60.00){\makebox(0,0)[ct]{\shortstack{\scriptsize{$2$}}}}
\put(112.00,70.00){\makebox(0,0)[ct]{\shortstack{\scriptsize{$2$}}}}
\put(132.00,88.00){\makebox(0,0)[lt]{\shortstack{\scriptsize{$2$}}}}
\put(6.00,110.00){\makebox(0,0)[rc]{$E$}}
\put(15.00,50.00){\vector(1,0){136.00}}
\put(146.00,46.00){\makebox(0,0)[ct]{$n_1-n_2$}}
\end{picture}
\end{center}
\caption{\it {
The spectrum of boundary states for $a_2^{(1)}$ Toda
theory with a solitonic boundary condition at coupling constant
$B\approx\frac{1}{2}$. The vertical axis is the boundary state
energy.}
\label{spectrum1}}
\end{figure}
\begin{figure}
\begin{center}
\epsfbox{spectrum.eps}
\end{center}
\caption{\it
As the coupling constant becomes smaller, the number of boundary
states grows. This plot shows the spectrum as in figure
\ref{spectrum1} but at coupling constant
$B\approx\frac{1}{10}$.
\label{spectrum2}}
\end{figure}
The amplitudes
for the reflection of the particles off these boundary states are
given explicitly in \eq{Khnm} for $a_2^{(1)}$, in eqs. \eqref{k11}
and \eqref{k21} for the first class of boundary conditions in
$a_4^{(1)}$ and eqs. \eqref{k12} and \eqref{k22} for the second
class. We have also begun the analysis for $a_3^{(1)}$ in section
\ref{a3}.
In these reflection amplitudes there are also a large number of
physical
strip poles which do not correspond to new bound states but have
an explanation in terms of a boundary generalisation \cite{dorey98}
of the Coleman-Thun mechanism \cite{cole}. The way in which this works
conveys the strong impression that there is some magic at
work which awaits a deeper explanation.
\section{Exact S-matrices and reflection amplitudes\label{sect:boot}}
This section gives a brief review of the properties of
particle S-matrices and reflection amplitudes and the consistency
equations which they are subject to. For an excellent presentation of
this material see \cite{corri96}.
The reason why the full S-matrices of affine Toda theories can be
determined exactly is that these theories possess
higher-spin symmetries.
A higher-spin symmetry implies that the sum of some
power of the momenta of the particles is conserved during scattering.
This allows only processes in which the incoming particles have
the same momenta as the outgoing ones. Furthermore
a higher-spin symmetry acts on the particles by shifting them
by amounts proportional to powers of their momenta.
This allows one to separate
the interaction regions as illustrated in figure
\ref{factorization}. This implies that any S-matrix factorizes
into a product of 2-particle S-matrices.
\begin{figure}[b]
\unitlength 1mm
\linethickness{0.4pt}
\begin{picture}(84.00,30.00)(-30,8)
\put(10.00,10.00){\line(1,1){20.00}}
\put(10.00,30.00){\line(1,-1){20.00}}
\put(69.00,10.00){\line(0,1){20.00}}
\put(37.00,10.00){\line(1,1){20.00}}
\put(64.00,10.00){\line(1,1){20.00}}
\put(37.00,30.00){\line(1,-1){20.00}}
\put(64.00,30.00){\line(1,-1){20.00}}
\put(52.00,10.00){\line(0,1){20.00}}
\put(20.00,10.00){\line(0,1){20.00}}
\put(33.00,20.00){\makebox(0,0)[cc]{$=$}}
\put(60.00,21.00){\makebox(0,0)[cc]{$=$}}
\put(20.00,20.00){\circle*{4.00}}
\put(47.00,20.00){\circle*{4.00}}
\put(74.00,20.00){\circle*{4.00}}
\put(52.00,25.00){\circle*{4.00}}
\put(52.00,15.00){\circle*{4.00}}
\put(69.00,25.00){\circle*{4.00}}
\put(69.00,15.00){\circle*{4.00}}
\end{picture}
\caption{\it Higher-spin symmetries allow one to shift the different
particle trajectories independently, thereby separating
multi-particle scattering processes into products of two-particle
scatterings.
\label{factorization}}
\end{figure}
If the particles live on a half-line, then in addition to
scattering with each other they are also going to reflect off the
boundary. Ideally one would hope that a multi-particle process
will now factorise into a succession of two-particle scatterings
and single-particle reflections off the boundary, as depicted in
figure~\ref{bfact}.
\begin{figure}
\unitlength 1.00mm
\linethickness{0.4pt}
\begin{picture}(121.42,36.00)(0,18)
\put(120.00,28.00){\line(-3,-2){15.00}}
\put(120.00,28.00){\line(-3,2){15.00}}
\put(120.00,42.00){\line(-2,-5){10.00}}
\put(85.00,28.00){\line(-2,5){10.00}}
\put(85.00,42.00){\line(-3,-2){15.00}}
\put(85.00,42.00){\line(-3,2){15.00}}
\put(50.00,35.00){\line(-3,-2){15.00}}
\put(50.00,35.00){\line(-3,2){15.00}}
\put(50.00,35.00){\circle*{4.00}}
\put(77.00,47.00){\circle*{2.83}}
\put(85.00,42.00){\circle*{2.83}}
\put(81.00,39.00){\circle*{2.83}}
\put(85.00,28.00){\circle*{2.83}}
\put(120.00,42.00){\circle*{2.83}}
\put(116.00,31.00){\circle*{2.83}}
\put(120.00,28.00){\circle*{2.83}}
\put(112.00,23.00){\circle*{2.83}}
\put(63.00,35.00){\makebox(0,0)[cc]{$=$}}
\put(99.00,35.00){\makebox(0,0)[cc]{$=$}}
\put(85.00,28.00){\line(-1,-2){5.00}}
\put(120.00,42.00){\line(-1,2){5.50}}
\put(50.00,35.00){\line(-1,-2){7.50}}
\put(50.00,35.00){\line(-1,2){7.50}}
\put(85.00,16.00){\rule{3pt}{37mm}}
\put(120.00,16.00){\rule{3pt}{37mm}}
\put(50.00,16.00){\rule{3pt}{37mm}}
\end{picture}
\caption{\it In the presence of an integrable boundary any
multi-particle process factorises into a product of two-particle
scatterings and single-particle reflections.
\label{bfact}}
\end{figure}
Of course all momentum-dependent translation
symmetries in the space direction will be broken by the boundary.
However one can choose the boundary condition in such a way that
it preserves those parts of the higher-spin-symmetries which
effect momentum-dependent translations in the time direction and
these suffice to separate the interaction regions. Thus in order to
determine the full S-matrix on the half-line we only need
the single-particle reflection matrices and the two-particle
S-matrices. The later will not be changed by the presence of the
boundary because
the particle scatterings could be shifted arbitrarily far away
from the boundary.
In affine Toda theory all particles are distinguished by their
masses and higher-spin charges and therefore the S-matrix has to
be diagonal.
\begin{equation}
\unitlength 1mm
\linethickness{0.4pt}
\begin{picture}(47.00,33.00)(20,0)
\put(25.00,5.00){\line(1,1){20.00}}
\put(25.00,25.00){\line(1,-1){20.00}}
\put(35.00,15.00){\circle*{2.00}}
\put(25.00,2.00){\makebox(0,0)[ct]{\scriptsize{$a$}}}
\put(45.00,2.00){\makebox(0,0)[ct]{\scriptsize{$b$}}}
\put(25.00,27.00){\makebox(0,0)[cb]{\scriptsize{$c$}}}
\put(45.00,27.00){\makebox(0,0)[cb]{\scriptsize{$d$}}}
\put(35.00,8.00){\makebox(0,0)[ct]{\scriptsize{$i(\th_1-\th_2)$}}}
\put(22.00,15.00){\makebox(0,0)[rc]{$S_{ab}^{cd}(\theta_1,\theta_2)\ =$}}
\put(47.00,15.00){\makebox(0,0)[lc]
{$=\ \delta_{bc}\,\delta_{ad}\,S_{ab}(\theta_1-\theta_2)$.}}
\put(35.00,12.50){\oval(4.00,3.00)[b]}
\end{picture}
\vspace{5mm}
\end{equation}
We have introduced the rapidity variable $\th$ which is
related to the momentum components by
\begin{equation}
E=m\cosh\th, ~~~~p=m\sinh\th,
\end{equation}
where $m$ is the mass of the particle.
We draw the worldline of a particle with rapidity $\theta$ at an
angle of $u=-i\theta$ to the vertical.
Depending on which of the higher-spin symmetries
remain unbroken by the boundary condition a particle has to be
reflected either into itself or into its antiparticle.
We will find that for the boundary conditions which we study
the former is the case and so the reflection matrices are diagonal
\begin{equation}
\unitlength 1.00mm
\linethickness{0.4pt}
\begin{picture}(45.00,25.00)(20,18)
\put(40.00,20.00){\rule{3pt}{20.00mm}}
\put(30.00,20.00){\line(1,1){10.00}}
\put(40.00,30.00){\line(-1,1){10.00}}
\put(40.00,30.00){\circle*{2.00}}
\put(38.50,26.50){\oval(3.00,3.00)[b]}
\put(37.00,21.00){\makebox(0,0)[cc]{$i\th$}}
\put(27.00,30.00){\makebox(0,0)[rc]{$K_a^b(\th)\ =$}}
\put(45.00,30.00){\makebox(0,0)[lc]{$=\ \delta_{ab}\,K_a(\th)$}}
\put(28.00,19.00){\makebox(0,0)[lt]{\scriptsize{$a$}}}
\put(28.00,41.00){\makebox(0,0)[lb]{\scriptsize{$a$}}}
\end{picture}
\end{equation}
The scattering and the reflection amplitudes have to be unitary,
real analytic and $2\pi i$-periodic. This implies that they
can be written as a
product of fundamental building blocks (we use the notation of
\cite{brade90} in which $h=n+1$ is the Coxeter number of $a_n^{(1)}$)
\begin{equation}\label{bracketre}
\biggl( x\biggr)\equiv\frac{\sin\left(\frac{\th}{2i}+\frac{\pi}{2h}x\right)}
{\sin\left(\frac{\th}{2i}-\frac{\pi}{2h}a\right)},
\end{equation}
\begin{equation}
S_{ab}(\th)=\prod_{x\in A_{ab}}\biggl( x\biggr)\,,~~~~~~
K_a(\th)=\prod_{x\in B_a}\biggl( x\biggr)\;.
\end{equation}
Each block $(x)$ has a pole at $\th=x\,i\pi/h$ and a zero at
$\th=-x\,i\pi/h$.
Thus the problem of calculating the full exact S-matrix has been reduced
to the task of finding the sets $A_{ab}$ and $B_a$ of numbers specifying the
locations of the poles and zeros in the two-particle scattering amplitudes
$S_{ab}$ and the reflection amplitudes $K_a$ respectively.
As an illustration we give the scattering amplitudes for $a_2^{(1)}$
affine Toda theory which we will meet again in section \ref{subsect:mixed2}.
This theory has two conjugate particles of equal mass with
scattering amplitudes
\begin{gather}
S_{11}=S_{22}=\biggl( -2+B\biggr)\,\biggl( -B\biggr)\,\biggl( 2\biggr)\,,\label{S11}\\
S_{12}=S_{21}=-\,\biggl( -3+B\biggr)\,\biggl( -1-B\biggr)\,\biggl( 1\biggr)\,.\label{S12}
\end{gather}
Here $B$ is the usual coupling
constant dependent function
\begin{equation}\label{B}
B(\beta) =\frac 1{2\pi} \frac{\beta^2}{1+\frac{\beta^2}{4\pi}}\;,
\end{equation}
The reflection amplitudes depend on which integrable boundary
condition we choose. For example for the solitonic boundary
condition dealt with in section \ref{subsect:mixed2} we have
\begin{equation}\label{K1}
K_1=K_2=\biggl( -2\biggr)\,\biggl( -\frac B2\biggr)\,\biggl( 2+\frac B2 \biggr)\,
\biggl( \frac 12 - \frac B2\biggr)\, \biggl(
\frac 32 - \frac B2 \biggr)\, \biggl( \frac 32 + \frac B2 \biggr)\, \biggl( \frac 52 +
\frac B2 \biggr)\;.
\end{equation}
Crossing symmetry relates different amplitudes to each other,
\begin{equation}
S_{ab}(\th)=S_{b\bar{a}}(i\pi-\th),~~~~~
K_a(\th)K_{\bar a}(\th+i\pi)=S_{aa}(2\th),
\end{equation}
where $\bar{a}$ denotes the antiparticle to $a$ and is given for
$a_n^{(1)}$ Toda theory by $\bar{a}=h-a$.
The second of these relations was first derived by Ghoshal and
Zamolodchikov \cite{ghosh94}. It can be checked that the example given
above satisfies these crossing relations by using the following properties
of the blocks:
\begin{equation}
\biggl( a\biggr)_{\th+i\pi}=-\biggl( a+h\biggr),~~~~
\biggl( a\biggr)_{2\th}=-\biggl(\frac{a}{2}\biggr)\,\biggl(\frac a2 +h\biggr),~~~~
\biggl( a\biggr)_{-\th}=-\biggl( a\biggr)^{-1}.
\end{equation}
By going from the Mandelstamm variable $s$ to the relative rapidity variable
$\th$ the physical sheet has been mapped to the physical strip
$0<Im(\th)<\pi$ (see for example \cite{Zam} for details).
Every pole on the physical strip has to have a physical
explanation. In the example above the amplitude $S_{11}$ has a
pole from the block $(2)$ at $\theta=2 i\pi/3$ which
corresponds to the propagation of a particle of type $2$ in the forward
channel, as depicted in figure~\ref{fig:pole} a).
\begin{figure}
\unitlength 0.50mm
\linethickness{0.4pt}
\begin{picture}(154.00,61.00)(-35,10)
\put(30.00,30.00){\line(2,1){20.00}}
\put(50.00,40.00){\line(2,-1){20.00}}
\put(50.00,40.00){\line(0,1){20.00}}
\put(30.00,70.00){\line(2,-1){20.00}}
\put(50.00,60.00){\line(2,1){20.00}}
\put(35.00,29.00){\makebox(0,0)[lt]{\scriptsize{$1$}}}
\put(65.00,29.00){\makebox(0,0)[rt]{\scriptsize{$1$}}}
\put(65.00,71.00){\makebox(0,0)[rb]{\scriptsize{$1$}}}
\put(35.00,71.00){\makebox(0,0)[lb]{\scriptsize{$1$}}}
\put(52.00,50.00){\makebox(0,0)[lc]{\scriptsize{$2$}}}
\put(50.00,34.00){\makebox(0,0)[cc]{\scriptsize{$\frac{2 i\pi}{3}$}}}
\put(150.00,30.00){\rule{3pt}{20mm}}
\put(150.00,42.00){\line(-5,-3){20.00}}
\put(150.00,58.00){\line(-5,3){20.00}}
\put(154.00,35.00){\makebox(0,0)[lc]{\scriptsize{vacuum}}}
\put(154.00,50.00){\makebox(0,0)[lc]{\scriptsize{excited boundary}}}
\put(154.00,65.00){\makebox(0,0)[lc]{\scriptsize{vacuum}}}
\put(135.00,29.00){\makebox(0,0)[lt]{\scriptsize{$1$}}}
\put(135.00,71.00){\makebox(0,0)[lb]{\scriptsize{$1$}}}
\put(50.00,20.00){\makebox(0,0)[cc]{a)}}
\put(150.00,20.00){\makebox(0,0)[cc]{b)}}
\end{picture}
\caption{\it a) Simple poles in a scattering amplitude may be due to
the propagation of a bound state. b) Simple poles in a reflection
amplitude may be due to the formation of an excited boundary
state.
\label{fig:pole}}
\end{figure}
A further consequence of the higher-spin symmetries and the
corresponding momentum-dependent translation of particle lines is
the bootstrap principle. It
expresses the scattering amplitudes for bound states in terms of
those of the constituent particles. For example the bound state
pole discussed above leads to the following bootstrap equation
which gives $S_{12}$ in terms of $S_{11}$,
\begin{equation}
\unitlength 0.6mm
\linethickness{0.4pt}
\begin{picture}(185.00,45.00)(0,10)
\put(20.00,20.00){\line(2,1){30.00}}
\put(50.00,35.00){\line(2,-1){30.00}}
\put(50.00,35.00){\line(0,1){16.00}}
\put(123.00,20.00){\line(2,1){30.00}}
\put(153.00,35.00){\line(2,-1){30.00}}
\put(153.00,35.00){\line(0,1){15.00}}
\put(117.00,24.00){\line(6,1){68.00}}
\put(14.00,38.00){\line(6,1){68.00}}
\put(160.00,31.00){\circle*{4.00}}
\put(138.00,27.00){\circle*{4.00}}
\put(50.00,44.00){\circle*{4.00}}
\put(20.00,17.00){\makebox(0,0)[ct]{\scriptsize{$1$}}}
\put(80.00,17.00){\makebox(0,0)[ct]{\scriptsize{$1$}}}
\put(123.00,16.00){\makebox(0,0)[ct]{\scriptsize{$1$}}}
\put(183.00,16.00){\makebox(0,0)[ct]{\scriptsize{$1$}}}
\put(50.00,53.00){\makebox(0,0)[cb]{\scriptsize{$2$}}}
\put(153.00,53.00){\makebox(0,0)[cb]{\scriptsize{$2$}}}
\put(115.00,24.00){\makebox(0,0)[rc]{\scriptsize{$2$}}}
\put(12.00,38.00){\makebox(0,0)[rc]{\scriptsize{$2$}}}
\put(56.00,41.00){\makebox(0,0)[lt]{\footnotesize{$S_{12}$}}}
\put(138.00,22.00){\makebox(0,0)[ct]{\footnotesize{$S_{11}$}}}
\put(160.00,26.00){\makebox(0,0)[ct]{\footnotesize{$S_{11}$}}}
\put(96.00,34.00){\makebox(0,0)[cc]{$=$}}
\end{picture}
\end{equation}
This provides an independent consistency check on the amplitude $S_{12}$
given in \eqref{S12} which was already determined by crossing
symmetry. In practice there is a large number of such bootstrap
consistency equations and these, together with the requirement
that any pole on the physical strip needs to have a physical
explanation, are strong enough to determine the S-matrices of all
affine Toda theories uniquely.
Similar relations exist for the reflection amplitudes.
Firstly there is a bootstrap equation relating the reflection
amplitudes of the different particles to each other, depicted in
figure~\ref{figbootstrap}. Secondly
every pole on the physical strip requires a physical
explanation. For reflection amplitudes the physical strip in the
rapidity plane is bounded by $Im(\th)\leq\pi/2$.
Our example in \eqref{K1} has two simple poles on the physical
strip, namely at $\th=i\frac{\pi}{3}\left(\frac 12-\frac B2\right)$
and at $\th=i\frac{\pi}{3}\left(\frac 32-\frac B2\right)$. These
correspond to the propagation of excited boundary states as
illustrated in figure~\ref{fig:pole} b).
The reflection amplitudes for reflection of these excited
boundaries is determined by the following bootstrap
equation,
\begin{equation}\label{bb}
\unitlength 0.5mm
\linethickness{0.4pt}
\begin{picture}(122.00,45.00)(10,8)
\put(50.00,10.00){\rule{3pt}{20.00mm}}
\put(50.00,30.00){\line(-4,-5){16.00}}
\put(120.00,25.00){\line(-3,-1){30.00}}
\put(120.00,25.00){\line(-3,1){30.00}}
\put(120.00,10.00){\rule{3pt}{20.00mm}}
\put(120.00,40.00){\line(-3,-4){22.50}}
\put(50.00,40.00){\line(-3,-1){30.00}}
\put(50.00,40.00){\line(-3,1){30.00}}
\put(50.00,40.00){\circle*{4.00}}
\put(111.00,28.00){\circle*{4.00}}
\put(120.00,25.00){\circle*{4.00}}
\put(105.00,20.00){\circle*{4.00}}
\put(70.00,30.00){\makebox(0,0)[cc]{$=$}}
\end{picture}
\end{equation}
The bootstrap equations would probably again be strong enough to
determine the reflection amplitudes uniquely if one knew the
spectrum of excited boundary states a priori. However this is not
easy to derive directly and therefore we took a different route,
to be described below.
Not only tree diagrams as those in figure~\ref{fig:pole} give rise to
poles on the physical strip. Rather any Feynman diagram in which
some internal lines are on-shell can produce poles. For example
the box diagram in figure~\ref{ct}a) will lead to a second order
pole if all the internal lines can be simultaneously on shell
(which does not happen in $a_2^{(1)}$ but in higher
Toda theories). The order of the pole is determined by the number
of lines on shell minus twice the number of loop integrations. The
boundary diagram in figure~\ref{ct}(b) has six internal lines and
two loop momenta and will thus generically lead to a second order
pole unless the reflection amplitude in the middle of the diagram
does itself have poles or zeros at this point. This boundary
generalisation of the Coleman-Thun mechanism \cite{cole} was discovered by
Dorey et.al. in \cite{dorey98}.
\begin{figure}
\begin{center}
\unitlength 0.7mm
\linethickness{0.4pt}
\begin{picture}(149.00,50.00)
\put(40.00,40.00){\line(1,0){20.00}}
\put(60.00,40.00){\line(0,-1){20.00}}
\put(60.00,20.00){\line(-1,0){20.00}}
\put(40.00,20.00){\line(0,1){20.00}}
\put(40.00,40.00){\line(-1,1){10.00}}
\put(60.00,40.00){\line(1,1){10.00}}
\put(60.00,20.00){\line(1,-1){10.00}}
\put(40.00,20.00){\line(-1,-1){10.00}}
\put(149.00,50.00){\line(0,-1){40.00}}
\put(149.00,40.00){\line(-2,-1){10.00}}
\put(139.00,35.00){\line(2,-1){10.00}}
\put(149.00,30.00){\line(-2,-1){10.00}}
\put(139.00,25.00){\line(2,-1){10.00}}
\put(139.00,35.00){\line(-4,1){14.00}}
\put(139.00,25.00){\line(-4,-1){14.00}}
\put(50.00,6.00){\makebox(0,0)[cc]{a)}}
\put(139.00,6.00){\makebox(0,0)[cc]{b)}}
\end{picture}
\caption{\it Diagrams with internal lines on shell lead to additional
poles on the physical strip.\label{ct}}
\end{center}
\end{figure}
Thus to check a conjectured amplitude one should draw all possible
diagrams which can have internal lines on shell, determine what
kind of poles they lead to, and then compare with the pole
structure of the conjectured amplitude. These checks have been
carefully performed for the scattering amplitudes of affine Toda
theory on the whole line. We will do the same for our reflection
amplitudes for $a_2^{(1)}$ and $a_4^{(1)}$ Toda theory in section
\ref{sect:poles} and will find that it works quite miraculously.
A similar analysis has been performed for the
reflection amplitudes in the boundary scaling Lee-Yang model
\cite{dorey98}.
\section{Reflection amplitudes for Neumann and $(++\cdots ++)$
boundary conditions\label{sect:an}}
\subsection{Reflection in $a_2^{(1)}$ ATFT revisited}
In a recent paper \cite{gande98b} one of us constructed reflection
amplitudes for $a_2^{(1)}$ ATFT. We briefly recall the main results of this
work here.
The construction was based on new solutions to the $a_2^{(1)}$
boundary Yang-Baxter equation (BYBE).
Some of these new solutions, the so-called $K$-matrices, were conjectured to
describe the reflection of the fundamental solitons in imaginary
coupling $a_2^{(1)}$ ATFT.
By comparison with semi-classical calculations \cite{deliu98} it was
possible to identify the two reflection matrices which correspond
to the Neumann and
the uniform (+++) boundary conditions.
Using the boundary bootstrap equations,
the reflection amplitudes for the breathers in
the theory were constructed.
Due to the identification of the lowest breathers in the imaginary
coupling theory with the fundamental particles in the real coupling
theory, it was then possible to find the reflection amplitudes for the
particles in the real coupling $a_2^{(1)}$ ATFT via a simple analytic
continuation.
This led to the following particle reflection amplitudes,
corresponding to the Neumann and the uniform (+++) boundary conditions
respectively,
\begin{eqnarray}
K^{N}_1(\th) = K^{N}_2(\th)&=&\biggl( -2\biggr)\,
\biggl( -\frac B2\biggr)\, \biggl( 2+\frac B2\biggr) \;, \\ \nonumber \\
K_1(\th) = K_2(\th)&=&\biggl(
-2\biggr)\, \biggl( \frac B2-1\biggr)\, \biggl( 3-\frac B2\biggr) \;, \label{realref1}
\end{eqnarray}
in which the brackets are defined as in (\ref{bracketre}) with $h=3$.
These amplitudes are in agreement with classical and semi-classical
calculations \cite{corri94b,kim95b,perki98}.
Given that the $S$-matrices in the bulk theory have long been known to
be self-dual under a weak--strong coupling duality, it was widely
expected that a duality like that can also be found in ATFT on a
half-line. However, the above reflection amplitudes are
not self-dual, but instead are dual to each other, i.e.
\begin{equation}
K^{N}_1(\th) \longleftrightarrow
K_1(\th) \;,
\hspace{2cm} (\mbox{as}\;\;\; \beta
\longleftrightarrow \frac{4\pi}{\beta})\;.
\end{equation}
This generalises the analogous
result in the sinh-Gordon theory in \cite{corri97}.
In the following we demonstrate how these results can be
generalised to the case of $a_n^{(1)}$ ATFTs.
Just as in the $a_2^{(1)}$ case, our construction of particle reflection
amplitudes begins with proposing a reflection matrix for the fundamental
solitons in imaginary coupling ATFT with uniform $(++\cdots ++)$
boundary condition.
\subsection{Notation}
We use the notation of \cite{gande96}.
Whenever we deal with solitons or breathers, we use the
following parametrisation of the rapidity:
\[
\mu = -i \frac{h\lambda}{2\pi}\th\;,
\]
and the bracket notation
\begin{equation}
\biggl( a \br_{\hs{-5pt} I}
\equiv \frac{\left[ a \right]} {\left[ -a \right]}\;,
\;\;\;\;\;\;\;\;\;\;
\left[ a \right] \equiv \sin\left(\frac{\pi}{h\lambda}(\mu +a)\right)
\;, \label{bracketim}
\end{equation}
in which $h=n+1$ is the Coxeter number of the affine Lie algebra
$a_n^{(1)}$, and $\lambda$ is related to the imaginary Toda
coupling constant $\beta$ in the following way:
\begin{equation}\label{lambda}
\lambda = -\frac{4\pi}{\beta^2} - 1\;.
\end{equation}
This is related to the function $B(\beta)$ in \eq{B} which is used
in real coupling Toda theory by $\lambda=-2/B$.
We introduced the subscript $I$ in order to distinguish this
block notation from the block notation of \eq{bracketre}
which we use when dealing with the fundamental particles.
\subsection{The diagonal $a_n^{(1)}$ soliton reflection matrices}
Although new non-diagonal solutions to the $a_n^{(1)}$ BYBE have been found
recently \cite{gande99}, it will turn out that here we only need the
diagonal solutions.
In these \mbox{$K$-matrices} all non-diagonal entries are zero, which
significantly simplifies all other equations, like the bootstrap,
crossing and unitarity conditions.
We know from the classical calculations in \cite{deliu98} that
solitons are always reflected into antisolitons by the boundaries
which we are considering\footnote{Also a soliton-preserving boundary
conditions exists \cite{deliu98b} but we do not consider it
here.}.
We are therefore only interested in those solutions of the BYBE in which
the solitons change into antisolitons, which means the $K$-matrices
are maps:
\begin{equation}
K_{A^{(a)}}(\mu) : V_a \longrightarrow V_{n+1-a}\;, \;\;\;\;\;\;\;\; (a = 1,2,...,n)\;.
\end{equation}
It is fairly straightforward to show that there is a solution to the
$a_n^{(1)}$ BYBE of the form
\begin{equation}
K_{A^{(1)}}(\mu) = \left( \begin{array}{cccccc}
1 & 0 & \dots & & & \\
0 & 1 & & & & \\
\vdots & & \ddots & & & \\
& & & & & \\
& & & & & 1 \end{array} \right)\, {\cal A}_1(\mu)\;,
\end{equation}
which is a $n\times n$ square matrix, which maps the solitons in the
first multiplet into their charge conjugate partners in the $n$th
multiplet. ${\cal A}_1(\mu)$ is an overall scalar factor, which is
determined by the requirements of boundary unitarity and boundary
crossing.
Similarly as in the $a_2^{(1)}$ case from \cite{gande98b} the two
conditions of boundary unitarity and boundary crossing lead to the
following equations for ${\cal A}_1(\mu)$:
\begin{eqnarray}\label{aeq}
{\cal A}_1(\mu)\, \overline{\cal A}_1(-\mu) &=& 1\;, \nonumber \\ \nonumber \\
\frac{{\cal A}_1(-\mu+\frac{n+1}4 \lambda)}{{\cal A}_1(\mu+\frac{n+1}4
\lambda)} &=& \frac{\overline{\cal A}_1(-\mu+\frac{n+1}4 \lambda)}{\overline{\cal
A}_1(\mu+\frac{n+1}4 \lambda)} = F_{1,1}(2\mu)\;,
\end{eqnarray}
in which $\overline{\cal A}_1 = {\cal A}_n$ is the overall scalar
factor of the charge conjugate reflection matrix, and
$F_{1,1}(\mu)$ is the scalar factor of the $a_n^{(1)}$ $S$-matrix
which was given in \cite{gande96}.
We can solve these equations in the same way as it was done in
\cite{gande98b}, and we obtain the minimal solution
\begin{eqnarray}\label{aa}
{\cal A}_1(\mu) = \overline{\cal A}_1(\mu) = \biggl( - \frac{n+1}4 \lambda \br_{\hs{-5pt} I}\,
\prod_{k=1}^{\infty}
\frac{G_{2k}(2\mu, - \frac 32 (n+1)\lambda +1 )\, G_{2k}(2\mu, -\frac 12
(n+1) \lambda)} {G_{2k}(2\mu, - \frac 32 n\lambda -\frac{\lambda}2 +1
)\, G_{2k}(2\mu, -\frac 12 n\lambda -\frac 32 \lambda)}\;,
\end{eqnarray}
in which
\begin{equation}
G_j(\mu,a) \equiv \frac{\Gamma(\mu+jh\lambda+a)} {\Gamma(-\mu+jh\lambda+a)}\;.
\end{equation}
There is the freedom to add certain CDD factors to this solution
as described in subsection \ref{scdd}.
We have checked that these soliton reflection amplitudes agree
semiclassically with the soliton time-delays calculated in
\cite{deliu98} for the uniform $(++\cdots++)$ boundary condition.
In principle it would now be possible to obtain the other soliton $K$-matrices
$K_{A^{(a)}}(\mu)$ for $(a=1,\dots,n)$ by using the bootstrap
equations. However, since we are mainly interested in the reflection
amplitudes for the breathers and subsequently for the particles in the
real coupling theory, we can first construct the reflection amplitudes
for the particle of type $1$ and then use the bootstrap equations
directly in the real coupling theory.
\subsection{The breather and particle reflection amplitudes}
The scattering amplitude for two solitons of type $1$ and type $n$ ($=\overline 1$)
has a pole at $\mu=\frac{n+1}2\lambda-1$ corresponding to the formation of the
lowest breather state $B_1^{(1)}$. This leads to the
breather bootstrap equations which were described in great detail in
the appendix of \cite{gande98b}\footnote{For the sine-Gordon model the breather
reflection amplitudes were obtained by Ghoshal in \cite{ghosh94b}.}.
Without going into further detail
we can write down the result for the breather reflection
amplitude,
\begin{eqnarray}\label{kb1}
K_{B_1^{(1)}}(\mu) &=& {\cal A}\left(\mu-\frac{n+1}4 \lambda+\frac 12\right)\,
F_{1,1}(2\mu)\, {\cal A}\left(\mu+\frac{n+1}4 \lambda-\frac 12\right) \nonumber \\
&=& \biggl( -\frac n2 \lambda \br_{\hs{-5pt} I}\, \biggl( -\frac 12(n+1) \lambda+\frac 12 \br_{\hs{-5pt} I}\,
\biggl( -\frac{\lambda}2-\frac 12 \br_{\hs{-5pt} I}\;. \label{KB1}
\end{eqnarray}
The fact that solitons reflect into antisolitons in the charge conjugate
multiplet ensures that the breather reflection is purely diagonal.
Since the breather $B_1^{(1)}(\mu)$ corresponds to the first particle
in real coupling $a_n^{(1)}$ ATFT, we obtain the reflection amplitude for
this particle through analytic continuation $\beta \longrightarrow
i\beta$ in (\ref{KB1}).
We find
\begin{equation}\label{con}
K_{B_1^{(1)}}(\mu) \longrightarrow K_1(\th) = \biggl( -n \biggr)\, \biggl( n+1-\frac B2
\biggr)\, \biggl( -1+\frac B2 \biggr)\;,
\end{equation}
in which we now use the block notation (\ref{bracketre}) and
the parameter $B=-2/\lambda$.
In analogy to the construction of the exact \mbox{$S$-matrices}
\cite{brade90}, it is possible to construct the reflection amplitudes for
all other particles from this lowest reflection amplitude. In $a_n^{(1)}$ ATFT
two particles, say $a$ and $b$, can fuse into a particle $a+b$ at the
rapidity difference $\th = i\pi\frac{a+b}{n+1}$ (if $a+b <n+1$). The
bootstrap principle for an integrable theory on the half-line states
that the reflection of two particles is independent of whether the
fusion into a third particle occurs before or after they reflect
off
the boundary. This leads to the following boundary bootstrap equation
\cite{fring94}
which is illustrated in figure~\ref{figbootstrap}:
\begin{equation}
K_{a+b}(\th) = K_b\left(\th-i\pi\frac{a}{n+1}\right)\,
S_{a,b}\left(2\th+i\pi\frac{b-a}{n+1}\right)\,
K_a\left(\th+i\pi\frac{b}{n+1}\right)\;, \label{boundbootab}
\end{equation}
in which $S_{a,b}(\th)$ is the real coupling $a_n^{(1)}$ affine Toda
$S$-matrix, which can be written in the form
\begin{equation}
S_{a,b}(\th) = \prod_{\stackrel{a-b+1}{\text{step }2}}^{a+b-1}
\biggl( p+1\biggr)\, \biggl( p-1 \biggr)\, \biggl(
-p-1+B \biggr)\, \biggl( -p+1-B \biggr)\;. \label{anSmatrix}
\end{equation}
We have chosen $a\leq b$.
\begin{figure}[ht]
\begin{center}
\begin{picture}(330,190)(0,0)
\put(120,10){\rule{3pt}{140pt}}
\put(120,80){\line(-1,-1){30}}
\put(90,50){\line(-2,-1){60}}
\put(90,50){\line(-1,-2){15}}
\put(120,80){\line(-1,1){30}}
\put(90,110){\line(-2,1){60}}
\put(90,110){\line(-1,2){15}}
\put(160,78){\line(1,0){13}}
\put(160,82){\line(1,0){13}}
\put(300,10){\rule{3pt}{140pt}}
\put(300,80){\line(-4,-3){80}}
\put(300,80){\line(-4,3){24}}
\put(300,50){\line(-1,-2){15}}
\put(300,50){\line(-1,2){24}}
\put(276,98){\line(-1,1){26}}
\put(250,124){\line(-2,1){35}}
\put(250,124){\line(-1,2){10}}
\put(24,11){\shortstack{\footnotesize{$a$}}}
\put(70,11){\shortstack{\footnotesize{$b$}}}
\put(213,11){\shortstack{\footnotesize{$a$}}}
\put(280,11){\shortstack{\footnotesize{$b$}}}
\put(90,68){\shortstack{\scriptsize{$a+b$}}}
\put(265,111){\shortstack{\scriptsize{$a+b$}}}
\put(277,81){\shortstack{\scriptsize{$b$}}}
\put(289,89){\shortstack{\scriptsize{$a$}}}
\put(289,55){\shortstack{\scriptsize{$b$}}}
\put(24,143){\shortstack{\footnotesize{$b$}}}
\put(70,143){\shortstack{\footnotesize{$a$}}}
\put(208,143){\shortstack{\footnotesize{$b$}}}
\put(238,146){\shortstack{\footnotesize{$a$}}}
\end{picture}
\end{center}
\caption{\it The boundary bootstrap relation}\label{figbootstrap}
\end{figure}
\vspace{0.2cm}
We can now obtain any $K_a(\th)$ from $K_1(\th)$ by applying
(\ref{boundbootab}) successively $a-1$ times:
\begin{equation}
K_a(\th) = K_1\(\th+i\pi\frac{a-1}{n+1}\)\, \prod_{c=1}^{a-1}
K_1\(\th-i\pi\frac{a+1-2c}{n+1}\)\,
S_{a-c,1}\(2\th+i\pi\frac{3c-a-1}{n+1}\)\;. \label{KSK}
\end{equation}
Using the formula (\ref{anSmatrix}) for the $S$-matrix
we can compute the right hand side of (\ref{KSK}) explicitly, and we
obtain
\begin{equation}
K_a(\th) = \prod_{c=1}^a \biggl( c-1 \biggr)\, \biggl( c-n-1 \biggr)\, \biggl( -c+\frac
B2 \biggr)\, \biggl( -c-n-\frac B2 \biggr)\;. \label{Ka}
\end{equation}
It is interesting to note that this result agrees with a
conjecture which was made for these
reflection amplitudes already 5 years ago in \cite{corri94}.
It is easy to check that these reflection amplitudes satisfy
the necessary boundary crossing relation
\begin{equation}
K_a(\th)K_{\bar a}(\th+i\pi)=S_{aa}(2\th).
\end{equation}
We also find that the amplitudes are invariant under
charge-conjugation,
\begin{equation}
K_a(\th) = K_{\bar a}(\th)\;.
\end{equation}
The classical limit ($B\to 0$) of $K_a(\th)$ is
\begin{equation}
K_a(\th) \longrightarrow -\biggl( -a \biggr)\, \biggl( a+n+1 \biggr)\;.
\end{equation}
This agrees with the classical result in formula (4.22) of \cite{bowco96b}.
In the sine--Gordon theory \cite{corri97} as well as the $a_2^{(1)}$ ATFT
\cite{gande98b} it was found
that the positive boundary condition and the Neumann boundary condition
are dual to each other under the weak--strong coupling duality $\beta
\longrightarrow \frac{4\pi}{\beta}$. Therefore, we conjecture that the dual
reflection amplitude to (\ref{Ka}) describes the reflection of $a_n^{(1)}$
affine Toda particles from a boundary with Neumann boundary
condition.
The dual of (\ref{Ka}) can simply be obtained by changing $B \to
2-B$,
\begin{equation}
K_a(\th) \longrightarrow K^N_a(\th) = \prod_{c=1}^a \biggl( c-1 \biggr)\, \biggl(
c-n-1 \biggr)\, \biggl( -c+1 - \frac B2 \biggr)\, \biggl( -c+n+1 +\frac B2 \biggr)\;.
\label{KNeum}
\end{equation}
First we check the
classical limit of these amplitudes and we find
\begin{equation}
K^N(\mu) \longrightarrow 1\;, \;\;\;\;\;\;\;\; (B \to 0)\;,
\end{equation}
which supports the conjecture that $K_a^N(\mu)$ are the reflection
amplitudes for the Neumann boundary condition.
In fact, here we can go one step further since Kim \cite{kim95b} has
computed the reflection amplitudes for the Neumann boundary condition up
to first order in the coupling constant.
If we expand $K^N(\mu)$ in terms of $\beta^2$ and compare the
${\cal O}(\beta^2)$ term with the perturbative result of \cite{kim95b}
we find that they are indeed identical.
\subsection{CDD pole ambiguities\label{scdd}}
The scalar factor in the soliton reflection
matrix is not uniquely fixed by the equations \eqref{aeq}.
Rather the minimal solution given in \eq{aa}
can be multiplied by any factor
$\sigma(\mu)$ which satisfies
\begin{eqnarray}
\sigma(\mu) \sigma(-\mu) &=& 1 \;, \nonumber \\
\sigma(-\mu +\frac{n+1}4 \lambda) &=& \sigma(\mu+\frac{n+1} 4 \lambda) \;.
\end{eqnarray}
Such ambiguities are often referred to as CDD pole ambiguities.
In order for the soliton bootstrap equations to be satisfied the
factors must obey the additional condition
\begin{equation}\label{ext}
\sigma(\mu) = \prod_{k=1}^n \sigma\left(\mu+\frac{n+1}2 \lambda - k\lambda \right)\;.
\end{equation}
These equations are solved by the following expressions
\begin{equation}\label{cdd}
\sigma_c(\mu) = - \biggl( \frac c2 \lambda \biggr)_I \, \biggl( \frac{n+1}2 \lambda -\frac c2
\lambda \biggr)_I\;,
\end{equation}
for $c= 1,2,\dots \leq \frac{n+1}2$ (actually $c$ can run from $1$ to
$n$ but that does not give any new factors since $\sigma_c(\mu) =
\sigma_{n+1-c}(\mu)$.)
The inverse of this, i.e. $\( \sigma_c(\mu) \) ^{-1} = - \biggl(
- \frac c2 \lambda \biggr) \, \biggl( \frac c2 \lambda - \frac{n+1}2 \lambda \biggr) $, is also
a solution of \eq{ext}.
The possible CDD factors are products of any of these
blocks.
After the bootstrap to the breather reflection amplitude and
analytic continuation to the particle amplitude a
soliton CDD factor $\sigma_c(\mu)$ would lead to a particle CDD factor
\begin{equation}
k_c(\mu) = \biggl( \frac{n+1}2 +c + \frac B2 \biggr)\, \biggl( \frac{n+1}2 -c - \frac
B2 \biggr)\, \biggl( - \frac{n+1}2 +c - \frac B2 \biggr)\,\biggl( -\frac{n+1}2 - c +
\frac B2 \biggr)\;.
\end{equation}
The classical limit ($B \to 0$) of these factors is equal to $1$.
We suspect that it will be
impossible to find a physical explanation of the extra poles
supplied by these CDD factors. Therefore we will from now on work
with the solution without CDD factors for which we will show the
consistency of the pole structure in section \ref{sect:poles}.
\section{The vacuum solutions for the solitonic boundary conditions
\label{sect:vacuum}}
We will now discuss the vacuum solutions to real
coupling $a_n^{(1)}$ affine Toda theory on the left half-line
with boundary condition of the form \eqref{bcond},
\begin{equation}\label{bcond2}
\left.\beta\,\partial_x{\boldsymbol \phi}+\sum_{i= 0}^n \bp_i\,{\boldsymbol \alpha}_i
e^{\beta{\boldsymbol \alpha}_i\cdot{\boldsymbol \phi}/2}\right|_{x= 0}= 0
\end{equation}
Two of these boundary conditions are rather special because their
vacuum solution is simply the constant solution
${\boldsymbol \phi}= 0$. The simplest is the Neumann condition which is obtained
if all the boundary parameters $\bp_i$ are equal to zero. The
other is the boundary condition which has all $\bp_i$ equal to
$+1$. We will refer to it as the uniform $(++\cdots++)$ boundary condition.
The boundary condition which has all $\bp_i$ equal to
$-1$ is also satisfied by the constant solution ${\boldsymbol \phi}=0$ but for
this boundary condition there exists another real solution with
the same energy.
In \cite{deliu98} we explained that if the $\bp_i$ are equal
to $\pm 1$ with at least one being $-1$ and if they satisfy the
constraint
\begin{equation}\label{bpcond}
\prod_{i= 0}^n\bp_i= 1,
\end{equation}
then the vacuum solution is
obtained by taking a solution describing a stationary soliton
on the left half-line and continuing it to real $\beta$.
The soliton has to have a topological charge ${\boldsymbol \lambda}$ determined
by
\begin{equation}\label{bptop}
(-1)^{{\boldsymbol \alpha}_i\cdot{\boldsymbol \lambda}}= \bp_i,~~~~~~\forall i= 1,\dots,n.
\end{equation}
We will refer to these boundary conditions as the solitonic boundary
conditions.
For details of how we put a stationary soliton in front of the
boundary we refer to \cite{deliu98}. It turns out that a soliton
with topological charge ${\boldsymbol \lambda}$ gives the same vacuum solution as
a soliton with topological charge $-{\boldsymbol \lambda}$, which is good
because both ${\boldsymbol \lambda}$ and $-{\boldsymbol \lambda}$ give the same boundary
condition if substituted into \eq{bptop}.
In this section we will
explain how to find this topological charge ${\boldsymbol \lambda}$ given
the boundary parameters $\bp_i$.
It is known that the topological charges of the single soliton
solutions all lie in fundamental representations of $a_n$.
A fundamental representation $\Lambda_i$ is the representation whose
highest weight is the fundamental weight ${\boldsymbol \lambda}_i$.
The fundamental weights ${\boldsymbol \lambda}_i$ are defined by their
property
\[
{\boldsymbol \lambda}_i\cdot {\boldsymbol \alpha}_j= \delta_{ij},~~~~~i,j= 1,\dots,n.
\]
Any ${\boldsymbol \lambda}$ can be written as a linear combination of the
fundamental weights, ${\boldsymbol \lambda}= \sum_{i= 1}^n a_i{\boldsymbol \lambda}_i$. It is
clear
that \eq{bptop} implies that the coefficient $a_i$ is even
if $\bp_i= 1$ and it is odd if $\bp_i= -1$. We will
show below that the unique positive weight
${\boldsymbol \lambda}$ which satisfies \eq{bptop} and which lies
in a fundamental representation is obtained by setting all
the even $a_i$ to zero and letting the odd $a_i$ alternate between
$+1$ and $-1$ (with the highest non-vanishing coefficient being
$+1$ so that the weight is positive).
Furthermore one can determine the particular
fundamental representation $\Lambda_a$ in which the weight ${\boldsymbol \lambda}$
lies by adding
the indices of the fundamental weights multiplied by their
coefficient in the expansion of ${\boldsymbol \lambda}$. We illustrate this rule by
a few examples:
\begin{align}
{\bf \bp}&= (-1,-1)&
{\boldsymbol \lambda}&= {\boldsymbol \lambda}_2-{\boldsymbol \lambda}_1
\in\Lambda_i&\text{with }&i= 2-1= 1,\\
{\bf \bp}&= (1,-1,1,-1)&
{\boldsymbol \lambda}&= {\boldsymbol \lambda}_4-{\boldsymbol \lambda}_2
\in\Lambda_i&\text{with }&i= 4-2= 2,\\
{\bf \bp}&= (-1,-1,1,1,-1)&
{\boldsymbol \lambda}&= {\boldsymbol \lambda}_5-{\boldsymbol \lambda}_2+{\boldsymbol \lambda}_1
\in\Lambda_i&\text{with }&i= 5-2+1= 4,
\end{align}
where ${\bf \bp}= (\bp_1,\bp_2,\dots,\bp_n)$. $\bp_0$ is not needed
because it is determined in terms of the other $\bp_i$'s by
the condition in \eqref{bpcond}, namely $\bp_0= \prod_{i= 1}^n\bp_i$.
Let us summarise the above rule for the relation between the
boundary parameters $\bp_i$ and the topological charge
$\lambda$ of the soliton in a theorem.
\begin{theorem}\label{t1}
Let $C_i$ be the parameters specifying a solitonic boundary.
Let $P= \{k_1,k_2,\dots,k_a\}$ be the ordered set of indices
(with $k_1<k_2<\dots<k_a$) for which the boundary parameters
are $-1$, i.e., $\bp_k= -1$ for $k\in P$.
If $P$ is not empty then there is a unique positive weight
${\boldsymbol \lambda}$ which satisfies
\begin{equation}\label{e46}
(-1)^{{\boldsymbol \alpha}_i\cdot{\boldsymbol \lambda}}= \bp_i,~~~~~~\forall i= 1,\dots,n,
\end{equation}
and lies in a fundamental representation. It is given by
\begin{equation}\label{e47}
{\boldsymbol \lambda}= \sum_{j= 1}^a(-1)^{a-j}{\boldsymbol \lambda}_{k_j},
\end{equation}
and lies in the $i$-th fundamental representation where
\begin{equation}\label{e48}
i= \sum_{j= 1}^a(-1)^{a-j}k_j.
\end{equation}
\end{theorem}
The rest of this section is devoted to the proof of this theorem
and can be skipped.
\vspace{3mm}
We needed one elementary fact
from the representation theory of simple complex Lie algebras:
\begin{theorem}\label{wt}
If ${\boldsymbol \lambda}$ is a weight of some irreducible representation
then ${\boldsymbol \lambda}-{\boldsymbol \alpha}_i$ is also a weight of the representation
iff ${\boldsymbol \lambda}\cdot{\boldsymbol \alpha}_i>0$. All weights of the representation
are obtained from the highest weight by repeatedly subtracting simple
roots in all possible
ways according to this rule.
\end{theorem}
We will use this to construct the weights of the fundamental
representations of $a_n$. It is convenient to introduce
an $n+1$ dimensional euclidean space with orthonormal basis vectors
$\e_i,~ i= 1,\dots,n+1$. We embed the $n$ dimensional root space of
$a_n$ into this space such that it is the
hyperplane perpendicular to the vector
$\e= \sum_{i= 1}^{n+1}\,\e_i$. An overcomplete and non-orthonormal
(but nevertheless useful)
basis for the root space is given by the
$n+1$ vectors
\begin{equation}
\er_i= \e_i-\frac{1}{n+1}\,\e,~~~~~i= 1,\cdots,n+1.
\end{equation}
It is overcomplete because $\er_{n+1}= -\sum_{i= 1}^n\er_i$.
The simple roots can be chosen to be
\begin{equation}
{\boldsymbol \alpha}_i= \e_i-\e_{i+1}= \er_i-\er_{i+1},~~~~~i= 1,\dots,n.
\end{equation}
The corresponding fundamental weights are
\begin{equation}\label{fw}
{\boldsymbol \lambda}_i= \sum_{a= 1}^i\er_a,~~~~~i= 1,\dots,n.
\end{equation}
We can now give the expressions for all the weights in the
fundamental representations.
\begin{proposition}
The weights of the $i$-th fundamental representation $\Lambda_i$
are all vectors that are given as a sum of $i$ different basis
vectors,
\begin{equation}
\Lambda_i= \left\{\left.\sum_{j= 1}^i\er_{k_j}\right|
k_j\neq k_{j'} \text{ for }j\neq j'\right\}.
\end{equation}
In other words, the weight vectors of $\Lambda_i$ are all vectors
with $i$ $1$'s and $n-i$ $0$'s when expanded in the $\er$ basis.
They are positive if the $n+1$st entry is 0.
The dimension of the $i$-th fundamental representation is
\begin{equation}
\rm{dim}\, \Lambda_i= \begin{pmatrix}n+1\\i\end{pmatrix}.
\end{equation}
\end{proposition}
\begin{proof}
Consider any weight vector ${\boldsymbol \lambda}$
whose entries are only $1$'s and $0$'s. The scalar
product of such a vector with a simple root ${\boldsymbol \alpha}_j$ is positive
only if the vector has a $1$ at the $j$-th position and a
$0$ at the $j+1$-st position. Then and only then will ${\boldsymbol \lambda}-{\boldsymbol \alpha}_j$
also be a weight, according to theorem \ref{wt}. The effect of
subtracting
${\boldsymbol \alpha}_j$ from ${\boldsymbol \lambda}$ will be to shift the $1$ at the $j$-th
position
one step to the right. Doing this repeatedly one can transform the
highest weight vector $\lambda_i= (1,\dots,1,0,\dots,0)$,
which has all the $i$ $1$'s to the left and the $0$'s
to the right, into any other vector with $i$ $1$'s. Furthermore one
will never obtain a weight vector with more or less than $i$ $1$'s.
The number of different vectors with $i$ $1$'s and the rest $0$'s is
equal to
the number of ways one can choose $i$ sites out of $n+1$ and this gives
the dimension of the $i$-th fundamental representation.
\end{proof}
If one inverts the expressions \eqref{fw}
for the fundamental weights
one arrives at the following expressions for the basis vectors in
terms of the fundamental weights
\begin{equation}
\er_1= {\boldsymbol \lambda}_1,~~~~
\er_i= {\boldsymbol \lambda}_i-{\boldsymbol \lambda}_{i-1},~~~i= 2,\dots,n,~~~~\er_{n+1}=
-{\boldsymbol \lambda}_n.
\end{equation}
Substituting these in the above proposition one obtains the following
corollary.
\begin{corollary}\label{cor}
The positive weights in the $i$-th fundamental representation are linear
combinations of fundamental weights,
$\lambda= \sum_{k= 1}^na_k{\boldsymbol \lambda}_k$,
with the following restrictions on the
coefficients:
\begin{enumerate}
\item The coefficient of
the largest occurring fundamental weight is $1$, the coefficient of
the next largest occurring fundamental weight is $-1$ and so on,
the non-zero coefficients alternating between $+1$ and $-1$.
\item The coefficients multiplied by the indices of
the corresponding fundamental weights sum up to $i$,
\begin{equation}
\sum_{k= 1}^n k\,a_k= i.
\end{equation}
\end{enumerate}
\end{corollary}
We are now in a position to prove our main theorem \ref{t1}.
The fact that a weight ${\boldsymbol \lambda}$ given by \eq{e47} lies in the $i$-th
fundamental representation with $i$ given by \eq{e48} is simply a
rewriting of Corollary \ref{cor}. It is obvious from
${\boldsymbol \lambda}_i\cdot{\boldsymbol \alpha}_j=\delta_{ij}$ that this ${\boldsymbol \lambda}$ satisfies
\eq{e46}. To
prove that ${\boldsymbol \lambda}$ is the unique positive weight in a
fundamental representation satisfying \eq{e46} we only have to
observe that
the total number of weights in the fundamental representations is
\begin{equation}
\sum_{i= 1}^n \text{dim}\Lambda_i= \sum_{i= 1}^n
\begin{pmatrix}n+1\\i\end{pmatrix}= 2^{n+1}-2
\end{equation}
and thus the number of positive weights is equal to $2^n-1$ which is
also equal to the number of boundary conditions which satisfy
\eq{bpcond} and don't have all $\bp_i$ equal to $1$.
\section{The reflection amplitudes for the solitonic boundary conditions
\label{sect:mixed}}
We can now utilise the result from the preceding section in order to
construct the reflection amplitudes for the solitonic boundary conditions.
We have seen that we should be able to obtain them
by putting a stationary soliton in front of the boundary. In the case
of an incoming breather this leads to an additional breather--soliton scattering
process before and after the reflection off the boundary as illustrated
in figure~\ref{figstatsol}. The new reflection amplitude is then
just the product of the two $S$-matrix elements and the reflection amplitude
for the uniform $(++\cdots ++)$ boundary.
\begin{figure}[ht]
\begin{center}
\begin{picture}(150,140)(-10,0)
\put(50,10){\rule{3pt}{120pt}}
\put(50,70){\line(-1,-1){45}}
\put(50,70){\line(-1,1){45}}
\put(35,15){\line(0,1){110}}
\put(65,67){\shortstack{ = $S_{B^{(a)}_1,A^{(b)}}(\mu)\, K_{B^{(a)}_1}(\mu)\,
S_{A^{(b)},B^{(a)}_1}(\mu)\;,$}}
\put(-5,15){\shortstack{\footnotesize{$B^{(a)}_1$}}}
\put(-2,120){\shortstack{\footnotesize{$B^{(a)}_1$}}}
\put(30,5){\shortstack{\footnotesize{$A^{(b)}_j$}}}
\end{picture}
\end{center}
\caption{\it A stationary soliton in front of the
boundary}\label{figstatsol}
\end{figure}
The scattering amplitudes of a soliton with a breather in imaginary
coupling $a_n^{(1)}$ ATFT was given in \cite{gande98}. There it was found
that this type of scattering process is independent of the
multiplet label $j$ of the soliton, and the scattering amplitudes were
found to be (using the bracket notation defined in \eq{bracketim})
\begin{equation}
S_{A^{(b)},B_1^{(a)}}(\mu) = \prod_{k=1}^a \frac{
\left[ \frac{\lambda}2 (a+b-\frac{n+1}2 -2k+2) +\frac 12 \right]\,
\left[ \frac{\lambda}2 (a-b-\frac{n+1}2 -2k) -\frac 12 \right]}
{\left[ \frac{\lambda}2 (a-b-\frac{n+1}2 -2k+2) +\frac 12 \right]\,
\left[ \frac{\lambda}2 (a+b-\frac{n+1}2 -2k) - \frac 12 \right]}\;,
\end{equation}
and
\begin{equation}
S_{B^{(a)}_1,A^{(b)}}(\mu) =
S_{A^{(n+1-b)},B_1^{(a)}}(\mu)\;. \label{BAsymm}
\end{equation}
Therefore, the reflection process in figure~\ref{figstatsol}
can be computed and we find
\begin{eqnarray}
\lefteqn{S_{B^{(a)}_1,A^{(b)}}(\mu)\, K_{B_1^{(a)}}(\mu)\,
S_{A^{(b)},B^{(a)}_1}(\mu) =} \hspace{1cm} \nonumber \\
&& = K_{B_1^{(a)}}(\mu)\,\prod_{k=1}^a
\frac{\biggl( \frac{\lambda}2 (a+b+\frac{n+1}2 -2k) -\frac 12 \br_{\hs{-5pt} I}\,
\biggl( \frac{\lambda}2 (b-a-\frac{n+1}2 +2k) +\frac 12 \br_{\hs{-5pt} I}}
{\biggl( \frac{\lambda}2 (a+b-\frac{n+1}2 -2k) -\frac 12 \br_{\hs{-5pt} I}\,
\biggl( \frac{\lambda}2 (b-a+\frac{n+1}2 +2k) + \frac 12 \br_{\hs{-5pt} I}} \;,
\label{addamplitude}
\end{eqnarray}
in which $K_{B_1^{(a)}}(\mu)$ is the amplitude describing the reflection
of the breather $B_1^{(a)}$ from the (++...+) boundary.
After analytic continuation the additional factor from the
two soliton--breather scattering amplitudes in \eq{addamplitude}
becomes
\begin{equation}
\kextra^{(b)}_a(\th)
= \prod_{k=1}^a \frac{\biggl( a+b+\frac{n+1}2 -2k +\frac B2 \biggr)\,
\biggl( b-a-\frac{n+1}2 +2k - \frac B2 \biggr)}
{\biggl( a+b-\frac{n+1}2 -2k + \frac B2 \biggr)\,
\biggl( b-a+\frac{n+1}2 +2k - \frac B2 \biggr)}\;.
\label{statsolfactor}
\end{equation}
We thus obtain the following reflection amplitudes for the particles in $a_n^{(1)}$ ATFT
with solitonic boundary conditions:
\begin{equation}
K^{(b)}_a(\th) = K_a(\th)\, \kextra^{(b)}_a(\th)\;, \label{Kmixed}
\end{equation}
in which $K_a(\th)$ is given by (\ref{Ka}).
Because of the symmetry relation (\ref{BAsymm}) we find that
\begin{equation}
K^{(b)}_a(\th) = K^{(n+1-b)}_a(\th)\,.
\end{equation}
This is needed for consistency because if a soliton of weight
${\boldsymbol \lambda}$ lies in representation $b$ then its antisoliton of
weight $-\lambda$ lies in the conjugate representation $n+1-b$ and
as we had mentioned in section \ref{sect:vacuum}, both of these solitons
give rise to the same vacuum solution and should therefore also
give the same particle reflection factors.
Something that wasn't obvious from the classical calculations is
the fact that all solitonic boundary conditions corresponding to
topological charges ${\boldsymbol \lambda}$ in the same fundamental
representation $V_b$ or its dual $V_{n+1-b}$
lead to the same reflection amplitudes.
There are only
$\left\lfloor\frac{n+1}{2}\right\rfloor$ classes of boundary conditions,
each corresponding to a pair of conjugate representations.
The classical vacuum solutions
look quite different for different members of these classes but their
quantum reflection amplitudes turn out to be the same.
It is again easy to check that the above reflection amplitudes
(\ref{Kmixed}) satisfy the necessary boundary crossing and
unitarity conditions.
We can also look at the classical limit of these amplitudes and we find
\begin{eqnarray}
\lefteqn{\kextra^{(a)}_b(\mu) \longrightarrow \prod_{k=1}^b \frac{\biggl(
a+b+\frac{n+1}2 -2k \biggr)\, \biggl( a-b-\frac{n+1}2 +2k \biggr)} {\biggl(
a+b-\frac{n+1}2 -2k \biggr)\,
\biggl( a-b+\frac{n+1}2 +2k \biggr)}} \hspace{0.5cm} \nonumber \\
&&= \biggl( a+b-\frac{n+1}2\biggr) \, \biggl( -a-b-\frac{n+1}2 \biggr) \, \biggl(
a-b+\frac{n+1}2\biggr) \, \biggl( b-a+\frac{n+1}2\biggr)\,,
\end{eqnarray}
which turns out to be the same as formula (4.21) in the paper
\cite{bowco96b}, whereas the dual of $\kextra^{(a)}_b(\mu)$ has again
a classical limit of $1$.
\section{Poles and boundary bound states
\label{sect:poles}}
In order to show the consistency of all reflection amplitudes we have
constructed, we need to examine their pole structure and
show that all physical strip poles can be explained in terms of
boundary bound states or in terms of some other processes which can
be regarded as analogues of the generalised Coleman-Thun mechanism in
the bulk theory \cite{dorey98}.
\subsection{Fixed poles\label{fixed}}
All $a_n^{(1)}$ affine Toda particle reflection amplitudes, as given in
(\ref{Ka}, \ref{KNeum}, \ref{Kmixed}), contain the same coupling constant
independent part and we can therefore discuss the pole structure in
this part collectively for all boundary conditions. Because
$K_a=K_{n+1-a}$ we only need to consider the case
$a\leq\frac{n+1}{2}$. Once
we have found the diagrams which explain the poles in $K_a$, the
poles in $K_{n+1-a}$ are explained by the charge conjugated
versions
of the same diagrams (i.e., we simply have to replace every
particle in a diagram by its antiparticle).
The term $\prod_{c=1}^a \biggl( c-1 \biggr)\, \biggl( c-n-1 \biggr)$ contains
poles on the physical strip at
\begin{equation}
\th = i\pi \frac{p}{n+1}\;, \hspace{1.5cm} p = 1,2,\dots,
a-1\;.
\end{equation}
The location of these poles does not depend on the
coupling constant.
They arise from the triangle diagram depicted in figure~\ref{figtriangle}
in which all lines are on-shell.
It is clear that this reflection process only exists for $p=1,2,\dots
a-1$ because otherwise particle $a-p$ does not exist.
\begin{figure}[ht]
\begin{center}
\begin{picture}(200,170)(0,0)
\put(120,10){\rule{3pt}{140pt}}
\put(85,115){\line(-1,2){15}}
\put(85,45){\line(0,1){70}}
\put(120,80){\line(-1,-1){35}}
\put(120,80){\line(-1,1){35}}
\put(85,45){\line(-1,-2){15}}
\put(65,7){\shortstack{\footnotesize{$a$}}}
\put(100,55){\shortstack{\footnotesize{$p$}}}
\put(103,101){\shortstack{\footnotesize{$p$}}}
\put(58,78){\shortstack{\footnotesize{$a-p$}}}
\put(65,148){\shortstack{\footnotesize{$a$}}}
\put(75,-3){\shortstack{\footnotesize{$\left(p\right)$}}}
\put(129.00,77.00){\makebox(0,0)[lc]{\shortstack{\footnotesize{$\left(a\right)$}}}}
\end{picture}
\end{center}
\caption{\it On-shell triangle diagram producing the pole at $\th=\frac{i\pi}{n+1}p$
in the reflection amplitude $K_a$ with $a\leq\frac{n+1}{2}$.\label{figtriangle}}
\end{figure}
The angles in the diagram are not drawn to scale. They are
determined by the requirement that the internal particle lines
should be on shell. The angles between the particle lines
and the vertical are equal to $-i$ times the rapidity of the
particles.
Thus, for example, to determine the angle
between the lines of particles $p$ and $a-p$ we only need to know
the relative rapidity at which particles $p$ and $a-p$ will fuse
to form the particle $a$. This is $\th=\frac{i\pi}{n+1}a$, as can
be read off from either the masses of the particles or from the
location of the corresponding pole in the S-matrix element
$S_{p,a-p}$. The angle is therefore equal to
$\frac{\pi}{n+1}a$. This is also the angle between the line of
particle $p$ and the boundary. In diagram \ref{figtriangle}, as in
all diagrams which follow, we have indicated the angle at which a
particle line meets the boundary by giving a number in parentheses
which states the angle as a multiple of $\frac{\pi}{n+1}$.
Similarly, we give the angle of the incoming line in the lower left
hand of the diagram. This gives the location of the pole produced
by the diagram.
Diagram \ref{figtriangle} contains three internal lines and
one loop and will therefore lead to a simple pole in the reflection
amplitude for particle $a$, provided that the reflection of
particle $p$ does not take place at a rapidity at which $K_p(\th)$
displays a pole or a zero. The rapidity of particle $p$ is
$\th=\frac{i\pi}{n+1}a$, as can be
read off from the angle given. Because $p<a$, $K_p$ does not have
any pole or zero at this location and therefore the diagram
does indeed lead to a simple pole.
This diagram explains all $\beta$ independent poles in all $a_n^{(1)}$ reflection
amplitudes. In the Neumann as well as the (++...+) boundary reflection
amplitudes there are no further poles. However, for the reflection
amplitudes (\ref{Kmixed}) for the solitonic boundaries
there are coupling constant dependent poles. Below we will study
the pole structure in detail for $a_2^{(1)}$ and $a_4^{(1)}$ Toda
theories. We will restrict our attention to a coupling constant in
the range $0<B<1$ (recall that in general $0\leq B\leq 2$) and we
will exclude any non-generic rational values at which special
cancellations might take place.
\subsection{$a_2^{(1)}$ ATFT\label{subsect:mixed2}}
In this subsection we have
\begin{equation}
\biggl( a \biggr) \equiv \frac{\sin(\frac{\th}{2i} +\frac{a\pi}{6})}
{\sin(\frac{\th}{2i} -\frac{a\pi}{6})}\;.
\end{equation}
In the case of $a_2^{(1)}$ there are three solitonic boundary conditions,
$(+ - -)$, $(- + -)$ and $(- - +)$.
Their vacuum solutions are given by placing the appropriate stationary
soliton from the vector representation in front of the boundary.
They all lead to the same reflection amplitude
\begin{equation}
K^{b_{0,0}}_1(\th) = K^{b_{0,0}}_2(\th) = K_1(\th) \biggl( \frac 12 - \frac B2
\biggr)\, \biggl(
\frac 32 - \frac B2 \biggr)\, \biggl( \frac 32 + \frac B2 \biggr)\, \biggl( \frac 52 +
\frac B2 \biggr)\;, \label{a2K}
\end{equation}
in which
\begin{equation}\label{kkk}
K_1(\th) =\biggl( -2 \biggr)\, \biggl( -1 +\frac B2 \biggr)\, \biggl( 3 -\frac B2
\biggr)\;= K_2(\th)\;.
\end{equation}
We have labelled the vacuum state by $b_{0,0}$ to distinguish it from
the excited boundary states $b_{n,m}$ to be derived
below.
In \cite{corri94} a calculation of the classical reflection
amplitude had lead to an expression for this reflection amplitude
containing an unknown function $C$ of the coupling constant. The conjecture
coincides with our result if one sets $C=-\frac{B}{2}$.
$K_1$ contains no poles in the physical strip. So there are two
physical strip poles in (\ref{a2K}), namely at
\begin{equation}
\th_l = \frac{i\pi}3\left(\frac 32 - \frac B2\right)\;, \hspace{1cm} \mbox{and}\;\;\;
\th_h = \frac{i\pi}3\left(\frac 12 - \frac B2\right)\;,
\end{equation}
where $\th_h$ only exists in the physical strip if $B\leq 1$.
We restrict our analysis to this case.
Let us begin with the pole $\th_h$ in $K_1^{b_{0,0}}$, which we expect to
correspond to a boundary bound
state (or excited boundary). The reflection amplitudes for the
reflection from this boundary bound state can easily be computed
using the boundary bootstrap equation \eq{bb}.
The boundary bound state formed by an incoming particle 1 with
rapidity $\th_h$ will be denoted by $b_{1,0}$.
The reflection amplitude for particle 1 scattering off boundary
$b_{1,0}$ is
\begin{eqnarray}
K_1^{b_{1,0}}(\th) &=& K_1(\th) \biggl( \frac 12 -\frac B2 \biggr)\, \biggl(
-\frac 12 -\frac B2
\biggr)\, \biggl( \frac 32 +\frac B2 \biggr)^2\nonumber\\
&& \hspace{30pt} \times \biggl( \frac 52 -\frac B2 \biggr)\, \biggl(
\frac 52 +\frac B2 \biggr)\,
\biggl( \frac 12 -\frac 32 B \biggr)\,
\biggl( -\frac 52
+\frac 32 B \biggr)\;.
\end{eqnarray}
We find that this reflection amplitude again contains physical strip
poles, some of which correspond to new boundary bound
states. Through repeated application of the boundary bootstrap
equation and by considering all possible physical strip poles, we
eventually obtain a whole array of boundary bound states, which
can be labelled by $b_{n,m}$, in which $n$ ($m$) is the number of
particles of type 1 (2) bound to the boundary.
In addition it turns out that $n$ or $m$ can be negative,
but only as long as $n+m \geq 0$. This ensures that there are no
boundary states with energies lower than the ground state
$b_{0,0}$.
The general reflection amplitude for a particle 1 scattering off a
boundary $b_{n,m}$ can be computed and we find
\begin{eqnarray}
K_1^{b_{n,m}}(\th) &=& K_1(\th)\, \biggl( \frac 12 - (2n+1)\frac B2
\biggr) \, \biggl( \frac 52 + (2m+1)\frac B2 \biggr) \nonumber \\
&& \hspace{31pt} \times \biggl( \frac 12 - (2n-1)\frac B2
\biggr) \, \biggl( \frac 52 + (2m-1)\frac B2 \biggr) \nonumber \\
&& \hspace{31pt} \times \biggl( \frac 32 + (2n-1)\frac B2
\biggr) \, \biggl( \frac 32 - (2m-1)\frac B2 \biggr) \nonumber \\
&& \hspace{31pt} \times \biggl( -\frac 52 + (2n+1)\frac B2
\biggr) \, \biggl( -\frac 12 - (2m+1)\frac B2 \biggr)\;. \label{Khnm}
\end{eqnarray}
It is easy to check that these reflection amplitudes all satisfy the
necessary boundary unitarity, crossing and bootstrap conditions.
The blocks in (\ref{Khnm}) have been arranged such that the second
block in each line is the crossing transform of the first block in
this line, i.e.\ each line is of the form $\biggl( a(n,m) \biggr) \, \biggl( h -
a(m,n) \biggr)$.
The reflection amplitudes for particle 2 can be obtained by
charge conjugation
\begin{equation}
K_2^{b_{n,m}}(\th) = K_1^{b_{m,n}}(\th)\;.
\end{equation}
In order to test the consistency of these reflection amplitudes, we now
have to explain all the poles in these new reflection amplitudes.
To do this we will need to know the possible particle fusings for
$a_2^{(1)}$. They are $1+1\rightarrow 2$ and $2+2\rightarrow 1$,
both occurring at a relative rapidity of $\th=2\frac{i\pi}{3}$.
First let us consider the case of $n+m > 0$. There are physical strip
poles in the first blocks on the first and second line and in the
second block on the third line of \eq{Khnm}, which can be explained by the
diagrams in figure~\ref{figa2pol1}.
\begin{figure}[ht]
\begin{picture}(480,190)(30,-10)
\put(120,10){\rule{3pt}{140pt}}
\put(120,55){\line(-1,-2){20}}
\put(120,105){\line(-1,2){20}}
\put(90,15){\shortstack{\scriptsize{$1$}}}
\put(95,150){\shortstack{\scriptsize{$1$}}}
\put(125,30){\shortstack{\footnotesize{$b_{n,m}$}}}
\put(125,77){\shortstack{\footnotesize{$b_{n+1,m}$}}}
\put(125,125){\shortstack{\footnotesize{$b_{n,m}$}}}
\put(280,10){\rule{3pt}{140pt}}
\put(280,110){\line(-1,-2){47}}
\put(280,50){\line(-1,2){47}}
\put(225,15){\shortstack{\scriptsize{$1$}}}
\put(230,150){\shortstack{\scriptsize{$1$}}}
\put(285,30){\shortstack{\footnotesize{$b_{n,m}$}}}
\put(285,77){\shortstack{\footnotesize{$b_{n-1,m}$}}}
\put(285,125){\shortstack{\footnotesize{$b_{n,m}$}}}
\put(440,10){\rule{3pt}{140pt}}
\put(440,26){\line(-1,2){18}}
\put(440,134){\line(-1,-2){18}}
\put(440,80){\line(-1,-1){18}}
\put(440,80){\line(-1,1){18}}
\put(422,62){\line(-2,1){60}}
\put(422,98){\line(-2,-1){60}}
\put(355,65){\shortstack{\scriptsize{$1$}}}
\put(355,92){\shortstack{\scriptsize{$1$}}}
\put(425,40){\shortstack{\scriptsize{$2$}}}
\put(426,118){\shortstack{\scriptsize{$2$}}}
\put(425,70){\shortstack{\scriptsize{$1$}}}
\put(425,85){\shortstack{\scriptsize{$1$}}}
\put(445,15){\shortstack{\footnotesize{$b_{n,m}$}}}
\put(445,55){\shortstack{\footnotesize{$b_{n,m-1}$}}}
\put(445,100){\shortstack{\footnotesize{$b_{n,m-1}$}}}
\put(445,140){\shortstack{\footnotesize{$b_{n,m}$}}}
\put(445,128){\shortstack{\tiny{$\(\frac 12-(2m-1)\frac B2\)$}}}
\put(445,72){\shortstack{\tiny{$\(\frac 12+(2m-1)\frac B2\)$}}}
\put(43,5){\shortstack{\scriptsize{$\(\frac 12 -(2n+1) \frac B2\)$}}}
\put(198,5){\shortstack{\scriptsize{$\(\frac 12 -(2n-1)\frac B2\)$}}}
\put(356,5){\shortstack{\scriptsize{$\(\frac 32 -(2m-1)\frac B2\)$}}}
\put(105,-15){\shortstack{ (a)}}
\put(260,-15){\shortstack{ (b)}}
\put(420,-15){\shortstack{ (c)}}
\end{picture}
\caption{\it Poles in $K_1^{b_{n,m}}$ with $n+m>0$.}\label{figa2pol1}
\end{figure}
Note that diagram \ref{figa2pol1}(c) only exists if $m > 0$.
This agrees with the fact that the corresponding pole in the
amplitude moves off the physical strip for $m\leq 0$.
There is an additional pole on the physical strip
in the case of $n\leq 0$ (but still $n+m >0$)
in the first block in the third line. This pole can be explained in
terms of the reflection process in figure~\ref{figa2pol2}.
\begin{figure}[ht]
\begin{center}
\begin{picture}(280,190)(0,-10)
\put(120,10){\rule{3pt}{140pt}}
\put(120,26){\line(-1,1){18}}
\put(120,134){\line(-1,-1){18}}
\put(120,80){\line(-1,-2){18}}
\put(120,80){\line(-1,2){18}}
\put(102,44){\line(-2,-1){60}}
\put(102,116){\line(-2,1){60}}
\put(35,10){\shortstack{\scriptsize{$1$}}}
\put(35,143){\shortstack{\scriptsize{$1$}}}
\put(108,28){\shortstack{\scriptsize{$1$}}}
\put(108,127){\shortstack{\scriptsize{$1$}}}
\put(106,61){\shortstack{\scriptsize{$2$}}}
\put(106,94){\shortstack{\scriptsize{$2$}}}
\put(125,15){\shortstack{\footnotesize{$b_{n,m}$}}}
\put(125,55){\shortstack{\footnotesize{$b_{n-1,m}$}}}
\put(125,100){\shortstack{\footnotesize{$b_{n-1,m}$}}}
\put(125,140){\shortstack{\footnotesize{$b_{n,m}$}}}
\put(45,0){\shortstack{\scriptsize{$\(\frac 32 +(2n-1)\frac B2\)$}}}
\put(125,130){\shortstack{\tiny{$\(\frac 12-(2n-1)\frac B2\)$}}}
\put(125,72){\shortstack{\tiny{$\(\frac 12+(2n-1)\frac B2\)$}}}
\end{picture}
\end{center}
\caption{\it Extra pole in $K_1^{b_{n,m}}$ for
$n \leq 0$.}\label{figa2pol2}
\end{figure}
It can be checked that the reflections in the centre of diagrams
\ref{figa2pol1}(c) and \ref{figa2pol2} occur at a zero in the
corresponding reflection amplitudes, and
therefore these diagrams yield simple poles, as expected.
Now we turn to the case of $n+m = 0$ which is slightly different.
In this case the second and fourth line in (\ref{Khnm}) cancel
each other and we get
\begin{eqnarray}
K_1^{b_{-n,n}}(\th) &=& K_1(\th)\, \biggl( \frac 12 + (2n-1)\frac B2
\biggr) \, \biggl( \frac 52 + (2n+1)\frac B2 \biggr) \nonumber \\
&& \hspace{31pt} \times \biggl( \frac 32 - (2n+1)\frac B2
\biggr) \, \biggl( \frac 32 - (2n-1)\frac B2 \biggr) \;. \label{Khn-n}
\end{eqnarray}
There are three possible poles in these amplitudes. The pole in $\biggl( \frac
12 + (2n-1)\frac B2 \biggr)$ corresponds to a process \ref{figa2pol1}a) in
which the
boundary $b_{-n+1,n}$ is created. The poles in $\biggl( \frac 32 -
(2n+1)\frac B2 \biggr)$ and in $\biggl( \frac 32 - (2n-1)\frac B2 \biggr)$
correspond to the two processes in figure~\ref{figa2pol3}(a) and
b) respectively.
\begin{figure}[ht]
\begin{picture}(480,190)(30,-10)
\put(120,10){\rule{3pt}{140pt}}
\put(120,55){\line(-1,-2){20}}
\put(120,105){\line(-1,2){20}}
\put(90,15){\shortstack{\scriptsize{$1$}}}
\put(95,150){\shortstack{\scriptsize{$1$}}}
\put(125,30){\shortstack{\footnotesize{$b_{-n,n}$}}}
\put(125,77){\shortstack{\footnotesize{$b_{-n-1,n+1}$}}}
\put(125,125){\shortstack{\footnotesize{$b_{-n,n}$}}}
\put(280,10){\rule{3pt}{140pt}}
\put(280,110){\line(-1,-2){47}}
\put(280,50){\line(-1,2){47}}
\put(225,15){\shortstack{\scriptsize{$1$}}}
\put(230,150){\shortstack{\scriptsize{$1$}}}
\put(285,30){\shortstack{\footnotesize{$b_{-n,n}$}}}
\put(285,77){\shortstack{\footnotesize{$b_{-n+1,n-1}$}}}
\put(285,125){\shortstack{\footnotesize{$b_{-n,n}$}}}
\put(440,10){\rule{3pt}{140pt}}
\put(440,26){\line(-1,2){18}}
\put(440,134){\line(-1,-2){18}}
\put(440,80){\line(-1,-1){18}}
\put(440,80){\line(-1,1){18}}
\put(422,62){\line(-2,1){60}}
\put(422,98){\line(-2,-1){60}}
\put(355,65){\shortstack{\scriptsize{$1$}}}
\put(355,92){\shortstack{\scriptsize{$1$}}}
\put(425,40){\shortstack{\scriptsize{$2$}}}
\put(426,118){\shortstack{\scriptsize{$2$}}}
\put(425,70){\shortstack{\scriptsize{$1$}}}
\put(425,85){\shortstack{\scriptsize{$1$}}}
\put(445,15){\shortstack{\footnotesize{$b_{n,-n}$}}}
\put(445,55){\shortstack{\footnotesize{$b_{n-1,-n+1}$}}}
\put(445,100){\shortstack{\footnotesize{$b_{n-1,-n+1}$}}}
\put(445,140){\shortstack{\footnotesize{$b_{n,-n}$}}}
\put(445,128){\shortstack{\tiny{$\(\frac 32-(2n-1)\frac B2\)$}}}
\put(445,72){\shortstack{\tiny{$\(-\frac 12+(2n-1)\frac B2\)$}}}
\put(43,5){\shortstack{\scriptsize{$\(\frac 32 -(2n+1) \frac B2\)$}}}
\put(198,5){\shortstack{\scriptsize{$\(\frac 32 -(2n-1)\frac B2\)$}}}
\put(356,5){\shortstack{\scriptsize{$\(\frac 52 -(2n-1)\frac B2\)$}}}
\put(105,-15){\shortstack{ (a)}}
\put(260,-15){\shortstack{ (b)}}
\put(420,-15){\shortstack{ (c)}}
\end{picture}
\caption{\it Poles in $K_1^{b_{-n,n}}$ and $K_1^{b_{n,-n}}$.}\label{figa2pol3}
\end{figure}
We also obtain
\begin{eqnarray}
K_1^{b_{n,-n}}(\th) &=& K_1(\th)\, \biggl( \frac 12 - (2n+1)\frac B2
\biggr) \, \biggl( \frac 52 - (2n-1)\frac B2 \biggr) \nonumber \\
&& \hspace{31pt} \times \biggl( \frac 32 + (2n-1)\frac B2
\biggr) \, \biggl( \frac 32 + (2n+1)\frac B2 \biggr) \;. \label{Kh-nn}
\end{eqnarray}
Here, the pole $\biggl( \frac 12 - (2n+1)\frac B2 \biggr)$ corresponds again to a
process of type \ref{figa2pol1}a) in which the boundary state
$b_{n+1,-n}$ is created.
Furthermore, if $\frac 2{2n-1} \leq B \leq \frac 3{2n-1}$ there
is a physical strip pole in $\biggl( \frac 52 - (2n-1) \frac B2 \biggr)$ which
can be explained by the process in diagram \ref{figa2pol3}(c).
It is easy to check that $K_1^{b_{n-1,-n+1}}(\th)$ displays a zero
at $\th=\frac{i\pi}{3}\left(-\frac 12+(2n-1)\frac B2\right)$
so that this process leads to a simple pole, as required.
For non-zero $B$ all the boundary bound state poles move outside of the
physical strip for large $n$ or $m$, and we therefore
obtain only a finite number of boundary bound states in the theory.
The following boundary bound states exist:
\begin{eqnarray}\label{boundsa2}
&b_{n,m}\;,&\mbox{for all integers} \hspace{5pt} n,m \hspace{5pt} \mbox{with}
\hspace{5pt} n+m > 0 \hspace{5pt} \mbox{and} \hspace{5pt}
-\frac 1{2B} - \frac 12 < n,m < \frac 1{2B}+\frac 12\;, \nonumber \\
&b_{n,-n} \hspace{4pt} \mbox{and} \hspace{4pt} b_{-n,n}\;, \hspace{5pt}
&\mbox{for all integers} \hspace{5pt} n \hspace{5pt} \mbox{with}
\hspace{5pt} 0\leq n < \frac 3{2B}+\frac 12\;. \nonumber
\end{eqnarray}
This spectrum is plotted in figures \ref{spectrum1} and
\ref{spectrum2} at two different values of $B$.
Note that a special phenomenon occurs whenever
$B=\frac{1}{2n}$ with $n$ integer. Then the states
$b_{-n-a,n+a}$ are degenerate with the states $b_{-n+a,n}$ for all
$a>0$ and similarly the states $b_{n+a,-n-a}$ are degenerate with
the states $b_{n,-n+a}$.
In the amplitude $K_1^{b_{-n,n}}$ given in \eq{Khn-n} the
blocks $\left(\frac 12+(2n-1)\frac B2\right)$ and
$\left(\frac 32-(2n+1)\frac B2\right)$ coincide at
$\left(1-\frac B2\right)$. One of them cancels against the block
$\left(-1+\frac B2\right)$ in $K_1$ (see \eq{kkk}).
This leaves
only a single pole which can receives a contribution both from
the propagation of
the boundary state $b_{-n+1,n}$ as in diagram \ref{figa2pol1}(a),
and from the propagation of the
boundary state $b_{-n-1,n+1}$ as in diagram \ref{figa2pol3}(a).
The degeneration in energy between these states can be observed
in figures~\ref{spectrum1} and \ref{spectrum2} which show the
spectrum at $B=\frac{1}{n}$ with $n=1$ or $n=5$ respectively.
The degeneration is displayed even more clearly by an animated plot
of the spectrum on this paper's web page at
{\tt http://www.mth.kcl.ac.uk/\~{}delius/reflection.html} which
plots the spin 2 conserved charge of the states against their
energy.
\subsection{$a_3^{(1)}$ ATFT\label{a3}}
For purely accidental reasons we have studied $a_4^{(1)}$ Toda
theory in detail instead of $a_3^{(1)}$.
However $a_3^{(1)}$ has a self-conjugate representation,
as do all $a_n^{(1)}$ Toda theories with $n$ odd.
This is a reason why one
might wish to look at $a_3^{(1)}$ Toda theory in detail.
The single stationary soliton
solutions for solitons in the self-conjugate representation
on the half-line have a slightly different form from the others.
Usually the
soliton solutions on the left half line are obtained by placing a mirror
antisoliton on the unphysical right half line. For the self-conjugate
soliton the soliton and its mirror antisoliton collapse into a single
soliton because they are both stationary and of the same type.
One still obtains real-valued non-singular vacuum
solutions after analytic continuation of these solutions to real coupling.
However one might expect that this difference in the classical theory
might lead to some new phenomenon in the reflection
amplitudes for the boundary conditions corresponding to a self-conjugate
representation.
In the $a_3^{(1)}$ Toda theory the
$(+ - + -)$, $(- + - +)$ and $(- - - -)$ boundary conditions
correspond to a soliton in the self-dual representation
$\Lambda_2$.
We will give some of the reflection amplitudes
for this case. The theory contains boundary states obtained by
binding $n_1$ particles of type $1$ and $n_3$ particles of type
$3$ to the boundary. The reflection amplitudes for particles
of type $1$ and $2$ reflecting off these boundaries are
\begin{align}\label{k312}
K_1^{n_1,n_3}(\th)\ = K_1&(\theta)
\biggl( 1 -(2n_1+1)\frac B2\biggr)\
\biggl( 3 +(2n_3+1) \frac B2\biggr)
&(a)\nonumber \\
&\times\biggl( 1 -(2n_1-1)\frac B2\biggr)\
\biggl( 3 +(2n_3-1) \frac B2\biggr)
&(b)\\
&\times\biggl( 1 +(2n_1-1)\frac B2\biggr)\
\biggl( 3 -(2n_3-1) \frac B2\biggr)
&(c)\nonumber \\
&\times\biggl(-1 -(2n_3+1) \frac B2\biggr)
\biggl(-3 -(2n_1+1)\frac B2\biggr)\
&(d)\nonumber
\end{align}
\begin{align}\label{k322}
K_2^{n_1,n_3}(\th)\ = K_2&(\theta)
\biggl( 2 -(2n_1-1)\frac B2\biggr)\
\biggl( 2 +(2n_3-1) \frac B2\biggr)
&(a)\nonumber \\
&\times\biggl( 2 -(2n_3-1) \frac B2\biggr)
\biggl( 2 +(2n_1-1)\frac B2\biggr)\
&(b)\\
&\times
\biggl( -(2n_1+1)\frac B2\biggr)\
\biggl( -4 +(2n_3+1) \frac B2\biggr)
&(c)\nonumber \\
&\times\biggl( -(2n_3+1)\frac B2\biggr)\
\biggl(-4 +(2n_1+1) \frac B2\biggr)
&(d)\nonumber
\end{align}
where
\begin{align}
K_1(\th)=&\biggl( -1+\frac B2\biggr)\;\biggl( -3\biggr)\;
\biggl( 4-\frac B2\biggr)\;,\\
K_2(\th)=&\biggl( -1+\frac B2\biggr)\;\biggl( -3\biggr)\;
\biggl( 4-\frac B2\biggr)\;
\biggl( 1\biggr)\;\biggl( 3-\frac B2\biggr)\;
\biggl( -2\biggr)\;\biggl(-2+\frac B2\biggr)\;.
\end{align}
The reflection amplitudes for particle $3$ are obtained by charge
conjugation,
\begin{equation}
K_3^{n_1,n_3}(\th)=K_1^{n_3,n_1}(\th).
\end{equation}
The pole in $K_1^{n_1,n_3}$ on line (a) is due to the propagation
of boundary state $b_{n_1+1,n_3}$ in the direct channel
(similar to diagram \ref{figa2pol1}(a)), that on line
(b) is due to the propagation of boundary state $b_{n_1-1,n_2}$ in
the crossed channel
(similar to diagram \ref{figa2pol1}(b)).
The pole on line $(c)$ is due to a diagram
similar to that in figure \ref{figa2pol2}.
Also for the poles in $K_2^{n_1,n_3}$
from the first block on line (a) and the first block on line (b)
one easily finds similar diagrams. This explains all poles for
$n_1$ and $n_2$ positive.
However if the pole in $K_1^{n_1,n_3}$ in line (b) is to be interpreted
as being due to the propagation of a state $b_{n_1-1,n_3}$ this means
that there will also have to be states with negative $n_1$ as long as
$n_1+n_2\geq 0$. The corresponding poles in $K_3^{n_1,0}$ also
require states with negative $n_3$. The reflection amplitude
$K_2^{n_1,n_3}$ develops additional poles in the first block
in line (c) if $n_1<0$ and from the first block in line (c) if
$n_3<0$. These poles indicate the existence of more
boundary bound states. We have not performed the bootstrap
calculations to determine the reflection amplitudes off these new
states. Thus it is an open question whether there is a consistent
explanation of all the poles in this case.
\subsection{$a_4^{(1)}$ ATFT}
This case we will treat in detail.
In this subsection we have
\begin{equation}
\biggl( a \biggr) \equiv \frac{\sin(\frac{\th}{2i} +\frac{a\pi}{10})}
{\sin(\frac{\th}{2i} -\frac{a\pi}{10})}\;.
\end{equation}
The fusion rules for the particles are that particles $a$ and $b$
can fuse to give particle $c$ if either $a+b=c$ or $a+b=c+5$.
The corresponding
fusion angles are
\begin{equation}
\theta_{ab}^c=\left\{\begin{array}{ll}
\frac{\pi}{5}(a+b)&\text{if } a+b=c\\
\frac{\pi}{5}(10-a-b)&\text{if } a+b=c+5
\end{array}\right.
\end{equation}
In the case of $a_4^{(1)}$ there are two classes of inequivalent
solitonic boundary conditions.
We will treat them in turns.
\subsubsection{The first class of boundary conditions}
The first class contains the boundary conditions
$(- + + + -)$, $(+ + + - -)$, $(+ + - - +)$,
$(- - + + +)$ and
$(+ - - + +)$. Their vacuum solutions are obtained by putting
solitons from the first or fourth fundamental representation in
front of the boundary.
The corresponding vacuum reflection amplitudes are
\begin{align}
K_1^{(1)0,0}(\th) = K_4^{(1)0,0}(\th)
&= K_1(\th) \biggl( \frac 52 - \frac B2 \biggr)\, \biggl(
\frac 52 + \frac B2 \biggr)\, \biggl( -\frac 12 - \frac B2 \biggr)\, \biggl( -\frac 92 +
\frac B2 \biggr)\;,\\
K_2^{(1)0,0}(\th) = K_3^{(1)0,0}(\th)
&= K_2(\th) \biggl( \frac 12 - \frac B2 \biggr)\, \biggl(
\frac 92 + \frac B2 \biggr)\, \biggl( \frac 32 - \frac B2 \biggr)\, \biggl( \frac 72 +
\frac B2 \biggr)\;.
\end{align}
$K_1$ and $K_2$ are given by \eq{Ka} and their poles on the physical
strip were explained in section \ref{fixed}.
The poles in $K_2^{(1)}(\th)$ and $K_3^{(1)}(\th)$ from the block
$\left( \frac 12 - \frac B2 \right)$ come from bound states of particles
$2$ and $3$ with the boundary respectively. Bootstrapping on these,
one discovers that there is a whole lattice of boundary bound
states $b_{n_2,n_3}$, where $n_i$ gives the number of particles of
type $i$ bound to the boundary. We allow $n_2$ or $n_3$ to become
negative, as long as the sum $n_2+n_3$ remains non-negative.
The reflection amplitudes off these
boundary bound states are found to be
\begin{align}
\label{k11}
K_1^{(1)n_2,n_3}(\th)\ = K_1&(\theta)
\biggl(\frac 32 -(2n_2-1)\frac B2\biggr)\
\biggl(\frac 72 +(2n_3-1) \frac B2\biggr)
&(a)\nonumber \\
&\times\biggl(\frac 52 -(2n_3-1)\frac B2\biggr)\
\biggl(\frac 52 +(2n_2-1) \frac B2\biggr)
&(b)\nonumber \\
&\times\biggl(-\frac 12 -(2n_2+1)\frac B2\biggr)\
\biggl(-\frac 92 +(2n_3+1) \frac B2\biggr)
&(c)\nonumber \\
&\times\biggl(-\frac 32 -(2n_3+1)\frac B2\biggr)\
\biggl(-\frac 72 +(2n_2+1) \frac B2\biggr)
&(d)
\end{align}
\begin{align}\label{k21}
K_2^{(1)n_2,n_3}(\th)\ = K_2(\theta)
&\biggl(\frac 12 -(2n_2+1)\frac B2\biggr)\
\biggl(\frac 92 +(2n_3+1) \frac B2\biggr)
&(a)\nonumber \\
&\times\biggl(\frac 12 -(2n_2-1)\frac B2\biggr)\
\biggl(\frac 92 +(2n_3-1) \frac B2\biggr)
&(b)\nonumber \\
&\times\biggl(\frac 32 -(2n_3-1)\frac B2\biggr)\
\biggl(\frac 72 +(2n_2-1) \frac B2\biggr)
&(c)\nonumber \\
&\times\biggl(\frac 32 +(2n_2-1)\frac B2\biggr)\
\biggl(\frac 72 -(2n_3-1) \frac B2\biggr)
&(d)\nonumber \\
&\times\biggl(\frac 52 -(2n_2-1)\frac B2\biggr)\
\biggl(\frac 52 +(2n_3-1) \frac B2\biggr)
&(e)\nonumber \\
&\times\biggl(-\frac 12 -(2n_3+1)\frac B2\biggr)\
\biggl(-\frac 92 +(2n_2+1) \frac B2\biggr)
&(f)\nonumber \\
&\times\biggl(-\frac 32 -(2n_2+1)\frac B2\biggr)\
\biggl(-\frac 72 +(2n_3+1) \frac B2\biggr)
&(g)\nonumber \\
&\times\biggl(-\frac 52 -(2n_3+1)\frac B2\biggr)\
\biggl(-\frac 52 +(2n_2+1) \frac B2\biggr)\;.
&(h)
\end{align}
The reflection amplitudes for particles $3$
and $4$ are obtained by charge conjugation,
\begin{equation}
K_3^{(1)n_2,n_3}(\theta)=K_2^{(1)n_3,n_2}(\theta),~~~~~
K_4^{(1)n_2,n_3}(\theta)=K_1^{(1)n_3,n_2}(\theta)
\end{equation}
and it is therefore sufficient if we explain the pole structure of
the reflection amplitudes for particles $1$ and $2$, the poles in the
reflection amplitudes for particles $3$ and $4$ are then explained by
the charge conjugated diagrams.
If $n_2+n_3$ equals to $0$ or $1$ then some
coincidences occur in the above products
of blocks. We will therefore first concentrate on the case
$n_2+n_3>1$.
$K_1^{(1)n_2,n_3}$ has physical strip poles in lines (a) and (b)
which are explained by the first two diagrams in figure
\ref{figa4pol1}.
If $n_3\leq 0$ the pole from the first factor on line (b)
is not on the physical sheet and diagram \ref{figa4pol1}(b) does not
exist.
If $n_2\leq 0$ then the second factor on line (b) is on the
physical sheet and is explained by the third diagram in figure
\ref{figa4pol1}.
It can be checked that the reflections in the middle of the three
processes in figure~\ref{figa4pol1} all occur at a zero in the
corresponding reflection amplitudes, and therefore these diagrams yield
simple poles as expected.
The physical strip poles in $K_2^{(1)n_2,n_3}$ can be explained by the
diagrams in figures \ref{figa4pol2}, \ref{figa4pol3} and
\ref{figa4pol4}.
We have drawn diagram \ref{figa4pol3}(d) for the case $n_2>0$. If
$n_2\leq 0$ then the diagram changes slightly to \ref{figa4pol4}(d).
Diagram \ref{figa4pol3}(e) only exists if $n_2>0$. For $n_2\leq 0$ the
corresponding pole moves off the physical strip.
There is an additional pole on the physical strip in the case of
$n_3<0$ in the second block in line (e) of $K_2^{(1)n_2,n_3}$.
This pole can be explained in terms of diagram \ref{figa4pol4}(e).
The diagrams \ref{figa4pol3} and
\ref{figa4pol4} all have
$13$ internal lines and $5$ loops. This would lead to a third
order poles. However it can be checked that the two reflection
factors appearing in every diagram contribute a zero, thereby
reducing the diagram to a simple pole. The scattering amplitudes
which appear in these diagrams do contribute neither a pole nor a
zero. Note that the diagrams \ref{figa4pol3}(e) and \ref{figa4pol4}(e) are
not as symmetric as we have drawn, rather the two reflections are taking
place at different angles, as indicated by the numbers in parentheses.
Next we turn to the case of $n_2=-n_3$ which is rather special.
In $K_1^{(1)n_2,n_3}$ there is a cancellation between line
(a) and line (d), leaving
\begin{align}
K_1^{(1)-n,n}(\th)\ =\ &\biggl(\frac 52 -(2n+1) \frac B2\biggr)
\biggl(\frac 52 -(2n-1)\frac B2\biggr)\
&(b)\\
&\times\biggl(-\frac 12 +(2n-1)\frac B2\biggr)\
\biggl(-\frac 92 +(2n+1) \frac B2\biggr)
&(c)\nonumber
\end{align}
In this case none of the diagrams in figure~\ref{figa4pol1}
exist, because they would all involve a state with $n_2+n_3<0$ which
we postulated not to exist. The poles in line (b) from
$\biggl( \frac 52 - (2n+1)\frac B2 \biggr)$ and
$\biggl( \frac 52 - (2n-1)\frac B2 \biggr)$
correspond to the first two processes in figure~\ref{figa4pol5}.
\begin{figure}[p]
\begin{picture}(480,190)(30,-15)
\put(120,10){\rule{3pt}{140pt}}
\put(120,26){\line(-1,2){18}}
\put(120,134){\line(-1,-2){18}}
\put(120,80){\line(-1,-1){18}}
\put(120,80){\line(-1,1){18}}
\put(102,62){\line(-2,1){60}}
\put(102,98){\line(-2,-1){60}}
\put(35,65){\shortstack{\scriptsize{$1$}}}
\put(35,92){\shortstack{\scriptsize{$1$}}}
\put(105,40){\shortstack{\scriptsize{$2$}}}
\put(106,118){\shortstack{\scriptsize{$2$}}}
\put(105,70){\shortstack{\scriptsize{$1$}}}
\put(105,85){\shortstack{\scriptsize{$1$}}}
\put(125,15){\shortstack{\footnotesize{$b_{n_2,n_3}$}}}
\put(125,55){\shortstack{\footnotesize{$b_{n_2-1,n_3}$}}}
\put(125,100){\shortstack{\footnotesize{$b_{n_2-1,n_3}$}}}
\put(125,140){\shortstack{\footnotesize{$b_{n_2,n_3}$}}}
\put(125,128){\shortstack{\tiny{$\(\frac 12-(2n_2-1)\frac B2\)$}}}
\put(125,74){\shortstack{\tiny{$\(\frac 12+(2n_2-1)\frac B2\)$}}}
\put(280,10){\rule{3pt}{140pt}}
\put(280,26){\line(-1,2){18}}
\put(280,134){\line(-1,-2){18}}
\put(280,80){\line(-1,-1){18}}
\put(280,80){\line(-1,1){18}}
\put(262,62){\line(-2,1){60}}
\put(262,98){\line(-2,-1){60}}
\put(195,65){\shortstack{\scriptsize{$1$}}}
\put(195,92){\shortstack{\scriptsize{$1$}}}
\put(265,40){\shortstack{\scriptsize{$3$}}}
\put(266,118){\shortstack{\scriptsize{$3$}}}
\put(265,70){\shortstack{\scriptsize{$2$}}}
\put(265,85){\shortstack{\scriptsize{$2$}}}
\put(285,15){\shortstack{\footnotesize{$b_{n_2,n_3}$}}}
\put(285,55){\shortstack{\footnotesize{$b_{n_2-1,n_3}$}}}
\put(285,100){\shortstack{\footnotesize{$b_{n_2-1,n_3}$}}}
\put(285,140){\shortstack{\footnotesize{$b_{n_2,n_3}$}}}
\put(285,128){\shortstack{\tiny{$\(\frac 12-(2n_3-1)\frac B2\)$}}}
\put(285,74){\shortstack{\tiny{$\(\frac 12+(2n_3-1)\frac B2\)$}}}
\put(440,10){\rule{3pt}{140pt}}
\put(440,26){\line(-1,1){18}}
\put(440,134){\line(-1,-1){18}}
\put(440,80){\line(-1,-2){18}}
\put(440,80){\line(-1,2){18}}
\put(422,44){\line(-2,-1){40}}
\put(422,116){\line(-2,1){40}}
\put(375,20){\shortstack{\scriptsize{$1$}}}
\put(375,135){\shortstack{\scriptsize{$1$}}}
\put(428,28){\shortstack{\scriptsize{$2$}}}
\put(428,127){\shortstack{\scriptsize{$2$}}}
\put(426,61){\shortstack{\scriptsize{$3$}}}
\put(426,94){\shortstack{\scriptsize{$3$}}}
\put(445,15){\shortstack{\footnotesize{$b_{n_2,n_3}$}}}
\put(445,55){\shortstack{\footnotesize{$b_{n_2-1,n_3}$}}}
\put(445,100){\shortstack{\footnotesize{$b_{n_2-1,n_3}$}}}
\put(445,140){\shortstack{\footnotesize{$b_{n_2,n_3}$}}}
\put(445,128){\shortstack{\tiny{$\(\frac 12-(2n_2-1)\frac B2\)$}}}
\put(445,74){\shortstack{\tiny{$\(\frac 12+(2n_2-1)\frac B2\)$}}}
\put(35,5){\shortstack{\scriptsize{$\left(\frac 32 -(2n_2-1)\frac B2\right)$}}}
\put(195,5){\shortstack{\scriptsize{$\left(\frac 52 -(2n_3-1)\frac B2\right)$}}}
\put(355,5){\shortstack{\scriptsize{$\left(\frac 52 +(2n_2-1)\frac B2\right)$}}}
\put(100,-15){\shortstack{ (a)}}
\put(260,-15){\shortstack{ (b)}}
\put(420,-15){\shortstack{ (b')}}
\end{picture}
\caption{\it Poles in $K_1^{(1)n_2,n_3}$ with $n_2+n_3>0$.}\label{figa4pol1}
\end{figure}
\begin{figure}[p]
\begin{picture}(480,190)(20,-15)
\put(120,10){\rule{3pt}{140pt}}
\put(120,55){\line(-1,-2){15}}
\put(120,105){\line(-1,2){20}}
\put(95,25){\shortstack{\scriptsize{$2$}}}
\put(95,150){\shortstack{\scriptsize{$2$}}}
\put(125,30){\shortstack{\footnotesize{$b_{n_2,n_3}$}}}
\put(125,77){\shortstack{\footnotesize{$b_{n_2+1,n_3}$}}}
\put(125,125){\shortstack{\footnotesize{$b_{n_2,n_3}$}}}
\put(280,10){\rule{3pt}{140pt}}
\put(280,110){\line(-1,-2){40}}
\put(280,50){\line(-1,2){47}}
\put(230,25){\shortstack{\scriptsize{$2$}}}
\put(230,150){\shortstack{\scriptsize{$2$}}}
\put(285,30){\shortstack{\footnotesize{$b_{n_2,n_3}$}}}
\put(285,77){\shortstack{\footnotesize{$b_{n_2-1,n_3}$}}}
\put(285,125){\shortstack{\footnotesize{$b_{n_2,n_3}$}}}
\put(440,10){\rule{3pt}{140pt}}
\put(440,26){\line(-1,2){18}}
\put(440,134){\line(-1,-2){18}}
\put(440,80){\line(-1,-1){18}}
\put(440,80){\line(-1,1){18}}
\put(422,62){\line(-2,1){60}}
\put(422,98){\line(-2,-1){60}}
\put(355,65){\shortstack{\scriptsize{$2$}}}
\put(355,92){\shortstack{\scriptsize{$2$}}}
\put(425,40){\shortstack{\scriptsize{$3$}}}
\put(426,118){\shortstack{\scriptsize{$3$}}}
\put(425,70){\shortstack{\scriptsize{$1$}}}
\put(425,85){\shortstack{\scriptsize{$1$}}}
\put(445,15){\shortstack{\footnotesize{$b_{n_2,n_3}$}}}
\put(445,55){\shortstack{\footnotesize{$b_{n_2,n_3-1}$}}}
\put(445,100){\shortstack{\footnotesize{$b_{n_2,n_3-1}$}}}
\put(445,140){\shortstack{\footnotesize{$b_{n_2,n_3}$}}}
\put(445,128){\shortstack{\tiny{$\(\frac 12-(2n_3-1)\frac B2\)$}}}
\put(445,74){\shortstack{\tiny{$\(\frac 32+(2n_3-1)\frac B2\)$}}}
\put(35,5){\shortstack{\scriptsize{$\left(\frac 12 -(2n_2+1)\frac B2\right)$}}}
\put(195,5){\shortstack{\scriptsize{$\left(\frac 12 -(2n_2-1)\frac B2\right)$}}}
\put(355,5){\shortstack{\scriptsize{$\left(\frac 32 -(2n_3-1)\frac B2\right)$}}}
\put(100,-15){\shortstack{ (a)}}
\put(260,-15){\shortstack{ (b)}}
\put(420,-15){\shortstack{ (c)}}
\end{picture}
\caption{\it Poles in $K_2^{(1)n_2,n_3}$ for $n_2+n_3>0$.}\label{figa4pol2}
\end{figure}
\begin{figure}[p]
\begin{picture}(480,240)(0,-15)
\put(120,5){\rule{3pt}{190pt}}
\put(120,26){\line(-1,2){18}}
\put(120,174){\line(-1,-2){18}}
\put(120,80){\line(-1,-1){18}}
\put(120,120){\line(-1,1){18}}
\put(120,80){\line(-1,1){48}}
\put(120,120){\line(-1,-1){48}}
\put(102,62){\line(-3,1){30}}
\put(102,138){\line(-3,-1){30}}
\put(72,72){\line(-4,-1){25}}
\put(72,128){\line(-4,1){25}}
\put(40,63){\shortstack{\scriptsize{$2$}}}
\put(40,134){\shortstack{\scriptsize{$2$}}}
\put(104,40){\shortstack{\scriptsize{$2$}}}
\put(106,158){\shortstack{\scriptsize{$2$}}}
\put(85,60){\shortstack{\scriptsize{$4$}}}
\put(85,135){\shortstack{\scriptsize{$1$}}}
\put(85,90){\shortstack{\scriptsize{$1$}}}
\put(85,106){\shortstack{\scriptsize{$3$}}}
\put(105,71){\shortstack{\scriptsize{$3$}}}
\put(106,125){\shortstack{\scriptsize{$1$}}}
\put(125,12){\shortstack{\footnotesize{$b_{n_2,n_3}$}}}
\put(125,95){\shortstack{\footnotesize{$b_{n_2-1,n_3}$}}}
\put(125,180){\shortstack{\footnotesize{$b_{n_2,n_3}$}}}
\put(125,167){\shortstack{\tiny{$\(\frac 12-(2n_2-1)\frac B2\)$}}}
\put(125,111){\shortstack{\tiny{$\(\frac 12+(2n_2-1)\frac B2\)$}}}
\put(125,72){\shortstack{\tiny{$\(\frac 12+(2n_2-1)\frac B2\)$}}}
\put(125,28){\shortstack{\tiny{$\(\frac 12-(2n_2-1)\frac B2\)$}}}
\put(320,5){\rule{3pt}{190pt}}
\put(320,23){\line(-1,2){20}}
\put(320,177){\line(-1,-2){20}}
\put(320,80){\line(-6,-5){20}}
\put(320,120){\line(-6,5){20}}
\put(320,80){\line(-6,5){42}}
\put(320,120){\line(-6,-5){42}}
\put(300,63){\line(-1,1){22}}
\put(300,137){\line(-1,-1){22}}
\put(278,85){\line(-4,1){25}}
\put(278,115){\line(-4,-1){25}}
\put(246,89){\shortstack{\scriptsize{$2$}}}
\put(246,106){\shortstack{\scriptsize{$2$}}}
\put(304,40){\shortstack{\scriptsize{$2$}}}
\put(305,158){\shortstack{\scriptsize{$2$}}}
\put(285,66){\shortstack{\scriptsize{$3$}}}
\put(285,128){\shortstack{\scriptsize{$1$}}}
\put(287,87){\shortstack{\scriptsize{$1$}}}
\put(286,109){\shortstack{\scriptsize{$4$}}}
\put(305,71){\shortstack{\scriptsize{$4$}}}
\put(306,123){\shortstack{\scriptsize{$1$}}}
\put(325,12){\shortstack{\footnotesize{$b_{n_2,n_3}$}}}
\put(325,95){\shortstack{\footnotesize{$b_{n_2-1,n_3}$}}}
\put(325,180){\shortstack{\footnotesize{$b_{n_2,n_3}$}}}
\put(325,167){\shortstack{\tiny{$\(\frac 12-(2n_2-1)\frac B2\)$}}}
\put(325,111){\shortstack{\tiny{$\(\frac 12+(2n_2-1)\frac B2\)$}}}
\put(325,72){\shortstack{\tiny{$\(\frac 32+(2n_2-1)\frac B2\)$}}}
\put(325,28){\shortstack{\tiny{$\(\frac 12-(2n_2-1)\frac B2\)$}}}
\put(35,5){\shortstack{\scriptsize{$\left(\frac 32 + (2n_2-1)\frac B2\right)$}}}
\put(235,5){\shortstack{\scriptsize{$\left(\frac 52 - (2n_2-1)\frac B2 \right)$}}}
\put(100,-15){\shortstack{ (d)}}
\put(300,-15){\shortstack{ (e)}}
\end{picture}
\caption{\it Poles in $K_2^{(1)n_2,n_3}$ if
$n_2>0$.}\label{figa4pol3}
\end{figure}
\begin{figure}[p]
\begin{picture}(480,240)(0,-15)
\put(120,5){\rule{3pt}{190pt}}
\put(120,25){\line(-1,1){15}}
\put(120,175){\line(-1,-1){15}}
\put(120,70){\line(-1,-2){15}}
\put(120,130){\line(-1,2){15}}
\put(120,70){\line(-1,2){51}}
\put(120,130){\line(-1,-2){51}}
\put(105,40){\line(-3,-1){36}}
\put(105,160){\line(-3,1){36}}
\put(69,28){\line(-1,-1){15}}
\put(69,172){\line(-1,1){15}}
\put(49,15){\shortstack{\scriptsize{$2$}}}
\put(47,188){\shortstack{\scriptsize{$2$}}}
\put(110,25){\shortstack{\scriptsize{$2$}}}
\put(110,169){\shortstack{\scriptsize{$2$}}}
\put(87,26){\shortstack{\scriptsize{$1$}}}
\put(87,168){\shortstack{\scriptsize{$4$}}}
\put(85,72){\shortstack{\scriptsize{$1$}}}
\put(85,122){\shortstack{\scriptsize{$3$}}}
\put(107,55){\shortstack{\scriptsize{$3$}}}
\put(107,140){\shortstack{\scriptsize{$1$}}}
\put(125,12){\shortstack{\footnotesize{$b_{n_2,n_3}$}}}
\put(125,95){\shortstack{\footnotesize{$b_{n_2-1,n_3}$}}}
\put(125,180){\shortstack{\footnotesize{$b_{n_2,n_3}$}}}
\put(125,167){\shortstack{\tiny{$\(\frac 12-(2n_2-1)\frac B2\)$}}}
\put(125,116){\shortstack{\tiny{$\(\frac 12+(2n_2-1)\frac B2\)$}}}
\put(125,64){\shortstack{\tiny{$\(\frac 32+(2n_2-1)\frac B2\)$}}}
\put(125,28){\shortstack{\tiny{$\(\frac 12-(2n_2-1)\frac B2\)$}}}
\put(320,10){\rule{3pt}{180pt}}
\put(320,26){\line(-1,2){18}}
\put(320,174){\line(-1,-2){18}}
\put(320,80){\line(-1,-1){18}}
\put(320,120){\line(-1,1){18}}
\put(320,80){\line(-1,1){48}}
\put(320,120){\line(-1,-1){48}}
\put(302,62){\line(-3,1){30}}
\put(302,138){\line(-3,-1){30}}
\put(272,72){\line(-4,-1){20}}
\put(272,128){\line(-4,1){20}}
\put(243,63){\shortstack{\scriptsize{$2$}}}
\put(243,133){\shortstack{\scriptsize{$2$}}}
\put(304,40){\shortstack{\scriptsize{$3$}}}
\put(306,158){\shortstack{\scriptsize{$3$}}}
\put(285,60){\shortstack{\scriptsize{$2$}}}
\put(285,135){\shortstack{\scriptsize{$4$}}}
\put(285,90){\shortstack{\scriptsize{$4$}}}
\put(284,106){\shortstack{\scriptsize{$1$}}}
\put(305,71){\shortstack{\scriptsize{$1$}}}
\put(306,125){\shortstack{\scriptsize{$4$}}}
\put(325,12){\shortstack{\footnotesize{$b_{n_2,n_3}$}}}
\put(325,95){\shortstack{\footnotesize{$b_{n_2,n_3-1}$}}}
\put(325,180){\shortstack{\footnotesize{$b_{n_2,n_3}$}}}
\put(325,167){\shortstack{\tiny{$\(\frac 12-(2n_3-1)\frac B2\)$}}}
\put(325,111){\shortstack{\tiny{$\(\frac 12+(2n_3-1)\frac B2\)$}}}
\put(325,72){\shortstack{\tiny{$\(\frac 32+(2n_3-1)\frac B2\)$}}}
\put(325,28){\shortstack{\tiny{$\(\frac 12-(2n_3-1)\frac B2\)$}}}
\put(35,5){\shortstack{\scriptsize{$\left(\frac 32 + (2n_2-1)\frac B2\right)$}}}
\put(235,5){\shortstack{\scriptsize{$\left(\frac 52 + (2n_3-1)\frac B2 \right)$}}}
\put(100,-15){\shortstack{ (d)}}
\put(300,-15){\shortstack{ (e')}}
\end{picture}
\caption{\it Pole (d) in $K_2^{(1)n_2,n_3}$ if $n_2\leq0$
and pole (e) if
$n_3<0$.}\label{figa4pol4}
\end{figure}
\begin{figure}[p]
\begin{picture}(380,190)(0,-15)
\put(120,10){\rule{3pt}{140pt}}
\put(120,55){\line(-1,-2){15}}
\put(120,105){\line(-1,2){20}}
\put(95,25){\shortstack{\scriptsize{$1$}}}
\put(95,150){\shortstack{\scriptsize{$1$}}}
\put(125,30){\shortstack{\footnotesize{$b_{-n,n}$}}}
\put(125,77){\shortstack{\footnotesize{$b_{-n-1,n+1}$}}}
\put(125,125){\shortstack{\footnotesize{$b_{-n,n}$}}}
\put(280,10){\rule{3pt}{140pt}}
\put(280,110){\line(-1,-2){42}}
\put(280,50){\line(-1,2){47}}
\put(230,25){\shortstack{\scriptsize{$1$}}}
\put(230,150){\shortstack{\scriptsize{$1$}}}
\put(285,30){\shortstack{\footnotesize{$b_{-n,n}$}}}
\put(285,77){\shortstack{\footnotesize{$b_{-n+1,n-1}$}}}
\put(285,125){\shortstack{\footnotesize{$b_{-n,n}$}}}
\put(45,5){\shortstack{\scriptsize{$\left(\frac 52 - (2n+1)\frac B2\right)$}}}
\put(205,5){\shortstack{\scriptsize{$\left(\frac 52 - (2n-1)\frac B2 \right)$}}}
\put(110,-15){\shortstack{(a)}}
\put(270,-15){\shortstack{(b)}}
\end{picture}
\caption{\it Poles in $K_1^{(1)-n,n}$.} \label{figa4pol5}
\end{figure}
\begin{figure}[p]
\begin{picture}(325.00,150.00)
(0,-15)
\put(120.00,10.00){\rule{3.00\unitlength}{140.00\unitlength}}
\put(125.00,22.00){\shortstack{\footnotesize{$b_{-n,n}$}}}
\put(125.00,94.00){\shortstack{\footnotesize{$b_{-n+1,n}$}}}
\put(125.00,135.00){\shortstack{\footnotesize{$b_{-n,n}$}}}
\put(125.00,39.00){\shortstack{\scriptsize{$\left(\frac 52 - (2n+1)\frac B2\right)$}}}
\put(125.00,67.00){\shortstack{\scriptsize{$\left(\frac 12 + (2n_2-1)\frac B2\right)$}}}
\put(320.00,10.00){\rule{3.00\unitlength}{140.00\unitlength}}
\put(320.00,26.00){\line(-1,1){18.00}}
\put(320.00,134.00){\line(-1,-1){18.00}}
\put(320.00,80.00){\line(-1,-2){18.00}}
\put(320.00,80.00){\line(-1,2){18.00}}
\put(302.00,44.00){\line(-2,-1){60.00}}
\put(302.00,116.00){\line(-2,1){60.00}}
\put(235.00,10.00){\shortstack{\scriptsize{$2$}}}
\put(235.00,143.00){\shortstack{\scriptsize{$2$}}}
\put(308.00,28.00){\shortstack{\scriptsize{$4$}}}
\put(308.00,127.00){\shortstack{\scriptsize{$4$}}}
\put(306.00,61.00){\shortstack{\scriptsize{$1$}}}
\put(306.00,94.00){\shortstack{\scriptsize{$1$}}}
\put(325.00,15.00){\shortstack{\footnotesize{$b_{-n,n}$}}}
\put(325.00,51.00){\shortstack{\footnotesize{$b_{-n-1,n+1}$}}}
\put(325.00,100.00){\shortstack{\footnotesize{$b_{-n-1,n+1}$}}}
\put(325.00,140.00){\shortstack{\footnotesize{$b_{-n,n}$}}}
\put(325.00,128.00){\shortstack{\scriptsize{$\left(\frac 52 + (2n+1)\frac B2\right)$}}}
\put(325.00,73.00){\shortstack{\scriptsize{$\left(\frac 12 - (2n+1)\frac B2\right)$}}}
\put(43.00,5.00){\shortstack{\scriptsize{$\left(\frac 32 - (2n+1)\frac B2\right)$}}}
\put(245.00,5.00){\shortstack{\scriptsize{$\left(\frac 32 - (2n+1)\frac B2 \right)$}}}
\put(110.00,-15.00){\shortstack{(a)}}
\put(310.00,-15.00){\shortstack{(b)}}
\put(120.00,80.00){\line(-1,-1){40.00}}
\put(80.00,40.00){\line(4,1){40.00}}
\put(80.00,40.00){\line(-2,-1){41.00}}
\put(120.00,80.00){\line(-1,1){40.00}}
\put(80.00,120.00){\line(4,-1){40.00}}
\put(80.00,120.00){\line(-2,1){41.00}}
\put(35.00,19.00){\makebox(0,0)[rc]{\shortstack{\scriptsize{$2$}}}}
\put(35.00,141.00){\makebox(0,0)[rc]{\shortstack{\scriptsize{$2$}}}}
\put(101.00,41.00){\makebox(0,0)[ct]{\shortstack{\scriptsize{$1$}}}}
\put(101.00,118.00){\makebox(0,0)[cb]{\shortstack{\scriptsize{$1$}}}}
\put(97.00,96.00){\makebox(0,0)[rt]{\shortstack{\scriptsize{$1$}}}}
\put(97.00,65.00){\makebox(0,0)[rb]{\shortstack{\scriptsize{$1$}}}}
\end{picture}
\caption{\it Extra pole in $K_2^{(1)-n,n}$ for a) $n\geq 0$ and b)
$n<0$.} \label{figa4pol5b}
\end{figure}
In $K_2^{(1)n_2,n_3}$ with $n_2=-n_3$ there are cancellations
between lines (b) and (f), (c) and (g) and (e) and (h), leaving only
\begin{align}
K_2^{(1)-n,n}(\th)\ =\ &\biggl(\frac 12 +(2n-1)\frac B2\biggr)\
\biggl(\frac 92 +(2n+1) \frac B2\biggr)
&(a)\nonumber \\
&\times\biggl(\frac 32 -(2n+1)\frac B2\biggr)\
\biggl(\frac 72 -(2n-1) \frac B2\biggr)
&(d)
\end{align}
The pole in line (a) still receives a contribution from the process
in figure~\ref{figa4pol2}(a). However
this pole does not satisfy the bootstrap equation which one would
expect on the basis of figure~\ref{figa4pol2}(a). The reason is that
there is now a second diagram
contributing to this pole, shown in figure~\ref{figa4pol5b}(a).
The processes in figures \ref{figa4pol2}(b) to
\ref{figa4pol4}(e) disappear, because they would involve boundary states
with $n_2+n_3<0$. The pole in line (d) from $\biggl(\frac 32 -(2n+1)\frac
B2\biggr)$ is explained by the diagram \ref{figa4pol5b}(d).
For a given $B$ with $0<B<1$ the following boundary bound
states exist
\begin{eqnarray}\label{boundsa4}
&b_{n_2,n_3}\;,&\mbox{for all integers} \hspace{3pt} n_2,n_3 \hspace{3pt} \mbox{with}
\hspace{3pt} n_2+n_3 > 0 \hspace{3pt} \mbox{and} \hspace{3pt}
-\frac 1{2B} - \frac 12 < n_2,n_3 < \frac 1{2B}+\frac 12\;, \nonumber \\
&b_{n,-n} \hspace{3pt} \mbox{and} \hspace{3pt} b_{-n,n}\;, \hspace{4pt}
&\mbox{for all integers} \hspace{3pt} n \hspace{3pt} \mbox{with}
\hspace{3pt} 0\leq n < \frac 5{2B}+\frac 12\;. \nonumber
\end{eqnarray}
Finally we consider the case where $n_2=-n_3+1$. In
$K_1^{(1)n_2,n_3}$ the two blocks in line (b) coincide and
give rise to a double pole. The reason for this is the fact that the
reflection process in the middle of diagram \ref{figa4pol1}(b) or
\ref{figa4pol1}(c) no longer provides a zero because it is a reflection
of a boundary with $n_2+n_3=0$.
Similarly in $K_2^{(1)n_2,n_3}$ the blocks in line (c) combine
with those in line (d) to form double poles, as do the two blocks in
line (e). Again this is because one of the reflection amplitudes in
each of the diagrams \ref{figa4pol2}(c), \ref{figa4pol3}(a), and
\ref{figa4pol3}(b) or \ref{figa4pol4}(b) ceases to supply a zero.
\subsubsection{The second class of boundary conditions}
The second class contains the boundary conditions $(- + + - +)$,
$(- + - + +)$, $(+ + - + -)$, $(- + - - -)$, $(+ - + + -)$, $(+ -
+ - +)$, $(- - + - -)$, $(- - - + -)$, $(- - - - +)$
and $(+ - - - -)$. Their vacuum solutions are obtained
by putting a soliton from the second or third fundamental
representation in front of the boundary.
The corresponding vacuum reflection amplitudes are
\begin{align}
K_1^{(2)}(\th) = K_4^{(2)}(\th)
= K_1(\th) &\biggl( \frac 12 - \frac B2 \biggr)\, \biggl(
\frac 92 + \frac B2 \biggr)\, \biggl( \frac 32 - \frac B2 \biggr)\, \biggl( \frac 72 +
\frac B2 \biggr)\;.\\
K_2^{(2)}(\th) = K_3^{(1)}(\th)
= K_2(\th) &\biggl( \frac 12 - \frac B2 \biggr)\, \biggl(
\frac 92 + \frac B2 \biggr)\, \biggl( \frac 32 - \frac B2 \biggr)\, \biggl( \frac 72 +
\frac B2 \biggr) \nonumber \\
&\times\biggl( \frac 52 -\frac B2 \biggr) \, \biggl( \frac 52 + \frac B2 \biggr)\,
\biggl( -\frac 12 - \frac B2 \biggr)\;.
\end{align}
All of these reflection amplitudes have a pole from a block
$\biggl( \frac 12 - \frac B2 \biggr)$, indicating that all particles can
bind to the boundary. Indeed we find a whole lattice of boundary bound
states $b_{n_1,n_2,n_3,n_4}$. For reasons to be explained below
this lattice is restricted from below by the three conditions
\begin{align}
n_1+n_2&\geq 0, &n_2+n_3&\geq 0, &n_3+n_4&\geq 0.
\end{align}
As previously the array is also bounded above because the poles at
which the bound states are created move off the physical strip.
Using the usual bootstrap equations, the reflection amplitudes off these
boundary bound states are found to be
\begin{align}\label{k12}
K_1^{(2)n_1,n_2,n_3,n_4}(\th)\ = K_1^{(1)n_2,n_3}&(\theta)
\biggl(\frac 12 -(2n_1+1)\frac B2\biggr)\
\biggl(\frac 92 +(2n_4+1) \frac B2\biggr)
&(e)\nonumber \\
&\times\biggl(\frac 12 -(2n_1-1)\frac B2\biggr)\
\biggl(\frac 92 +(2n_4-1) \frac B2\biggr)
&(f)\\
&\times\biggl(\frac 32 +(2n_1-1)\frac B2\biggr)\
\biggl(\frac 72 -(2n_4-1) \frac B2\biggr)
&(g)\nonumber \\
&\times\biggl(-\frac 52 -(2n_4+1)\frac B2\biggr)\
\biggl(-\frac 52 +(2n_1+1) \frac B2\biggr)
&(h)\nonumber
\end{align}
\begin{align}\label{k22}
K_2^{(2)n_1,n_2,n_3,n_4}(\th)\ = K_2^{(1)n_2,n_3}&(\theta)
\biggl(\frac 32 -(2n_1-1)\frac B2\biggr)\
\biggl(\frac 72 +(2n_4-1) \frac B2\biggr)
&(i)\nonumber \\
&\times\biggl(\frac 52 -(2n_4-1)\frac B2\biggr)\
\biggl(\frac 52 +(2n_1-1) \frac B2\biggr)
&(j)\\
&\times\biggl(-\frac 12 -(2n_1+1)\frac B2\biggr)\
\biggl(-\frac 92 +(2n_4+1) \frac B2\biggr)
&(k)\nonumber \\
&\times\biggl(-\frac 32 -(2n_4+1)\frac B2\biggr)\
\biggl(-\frac 72 +(2n_1+1) \frac B2\biggr)
&(l)\nonumber
\end{align}
\begin{figure}[p]
\begin{picture}(480,190)(20,-15)
\put(120,10){\rule{3pt}{140pt}}
\put(120,55){\line(-1,-2){15}}
\put(120,105){\line(-1,2){20}}
\put(95,25){\shortstack{\scriptsize{$1$}}}
\put(95,150){\shortstack{\scriptsize{$1$}}}
\put(125,30){\shortstack{\footnotesize{$b_{n_1,n_2,n_3,n_4}$}}}
\put(125,77){\shortstack{\footnotesize{$b_{n_1+1,n_2,n_3,n_4}$}}}
\put(125,125){\shortstack{\footnotesize{$b_{n_1,n_2,n_3,n_4}$}}}
\put(280,10){\rule{3pt}{140pt}}
\put(280,110){\line(-1,-2){42}}
\put(280,50){\line(-1,2){47}}
\put(230,25){\shortstack{\scriptsize{$1$}}}
\put(230,150){\shortstack{\scriptsize{$1$}}}
\put(285,30){\shortstack{\footnotesize{$b_{n_1,n_2,n_3,n_4}$}}}
\put(285,77){\shortstack{\footnotesize{$b_{n_1-1,n_2,n_3,n_4}$}}}
\put(285,125){\shortstack{\footnotesize{$b_{n_1,n_2,n_3,n_4}$}}}
\put(440,10){\rule{3pt}{140pt}}
\put(440,26){\line(-1,1){18}}
\put(440,134){\line(-1,-1){18}}
\put(440,80){\line(-1,-2){18}}
\put(440,80){\line(-1,2){18}}
\put(422,44){\line(-2,-1){45}}
\put(422,116){\line(-2,1){60}}
\put(370,20){\shortstack{\scriptsize{$1$}}}
\put(355,143){\shortstack{\scriptsize{$1$}}}
\put(428,28){\shortstack{\scriptsize{$1$}}}
\put(428,127){\shortstack{\scriptsize{$1$}}}
\put(426,61){\shortstack{\scriptsize{$2$}}}
\put(426,94){\shortstack{\scriptsize{$2$}}}
\put(445,15){\shortstack{\footnotesize{$b_{n_1,n_2,n_3,n_4}$}}}
\put(445,55){\shortstack{\footnotesize{$b_{n_1-1,n_2,n_3,n_4}$}}}
\put(445,100){\shortstack{\footnotesize{$b_{n_1-1,n_2,n_3,n_4}$}}}
\put(445,140){\shortstack{\footnotesize{$b_{n_1,n_2,n_3,n_4}$}}}
\put(445,128){\shortstack{\tiny{$\(\frac 12-(2n_1-1)\frac B2\)$}}}
\put(445,74){\shortstack{\tiny{$\(\frac 12+(2n_1-1)\frac B2\)$}}}
\put(35,5){\shortstack{\scriptsize{$\left(\frac 12 - (2n_1+1)\frac B2\right)$}}}
\put(195,5){\shortstack{\scriptsize{$\left(\frac 12 - (2n_1-1)\frac B2 \right)$}}}
\put(355,5){\shortstack{\scriptsize{$\left(\frac 32+ (2n_1+1)\frac B2 \right)$}}}
\put(110,-15){\shortstack{(e)}}
\put(250,-15){\shortstack{(f)}}
\put(410,-15){\shortstack{(g)}}
\end{picture}
\caption{\it Poles in $K_1^{(2)n_1,n_2,n_3,n_4}$}\label{figa4pol6}
\end{figure}
\begin{figure}[p]
\begin{picture}(480,190)(0,-15)
\put(120,10){\rule{3pt}{140pt}}
\put(120,30){\line(-3,4){50}}
\put(120,130){\line(-3,-4){50}}
\put(120,80){\line(-3,-1){50}}
\put(120,80){\line(-3,1){50}}
\put(70,63){\line(-1,-1){25}}
\put(70,96){\line(-1,1){25}}
\put(37,35){\shortstack{\scriptsize{$2$}}}
\put(37,120){\shortstack{\scriptsize{$2$}}}
\put(100,44){\shortstack{\scriptsize{$1$}}}
\put(102,115){\shortstack{\scriptsize{$1$}}}
\put(105,68){\shortstack{\scriptsize{$1$}}}
\put(105,87){\shortstack{\scriptsize{$1$}}}
\put(125,15){\shortstack{\footnotesize{$b_{n_1,n_2,n_3,n_4}$}}}
\put(125,55){\shortstack{\footnotesize{$b_{n_1-1,n_2,n_3,n_4}$}}}
\put(125,100){\shortstack{\footnotesize{$b_{n_1-1,n_2,n_3,n_4}$}}}
\put(125,140){\shortstack{\footnotesize{$b_{n_1,n_2,n_3,n_4}$}}}
\put(125,70){\shortstack{\tiny{$\(\frac 52-(2n_1-1)\frac B2\)$}}}
\put(125,125){\shortstack{\tiny{$\(\frac 12-(2n_1-1)\frac B2\)$}}}
\put(320,10){\rule{3pt}{140pt}}
\put(320,26){\line(-1,2){18}}
\put(320,134){\line(-1,-2){18}}
\put(320,80){\line(-1,-1){18}}
\put(320,80){\line(-1,1){18}}
\put(302,62){\line(-2,1){60}}
\put(302,98){\line(-2,-1){60}}
\put(235,65){\shortstack{\scriptsize{$2$}}}
\put(235,92){\shortstack{\scriptsize{$2$}}}
\put(305,40){\shortstack{\scriptsize{$4$}}}
\put(306,118){\shortstack{\scriptsize{$4$}}}
\put(305,70){\shortstack{\scriptsize{$2$}}}
\put(305,85){\shortstack{\scriptsize{$2$}}}
\put(325,15){\shortstack{\footnotesize{$b_{n_1,n_2,n_3,n_4}$}}}
\put(325,55){\shortstack{\footnotesize{$b_{n_1,n_2,n_3,n_4-1}$}}}
\put(325,100){\shortstack{\footnotesize{$b_{n_1,n_2,n_3,n_4-1}$}}}
\put(325,140){\shortstack{\footnotesize{$b_{n_1,n_2,n_3,n_4}$}}}
\put(325,128){\shortstack{\tiny{$\(\frac 12-(2n_4-1)\frac B2\)$}}}
\put(325,74){\shortstack{\tiny{$\(\frac 32+(2n_4-1)\frac B2\)$}}}
\put(45,5){\shortstack{\scriptsize{$\left(\frac 32 - (2n_1-1)\frac B2\right)$}}}
\put(245,5){\shortstack{\scriptsize{$\left(\frac 52 - (2n_4-1)\frac B2 \right)$}}}
\put(110,-15){\shortstack{(i)}}
\put(310,-15){\shortstack{(j)}}
\end{picture}
\caption{\it Poles in $K_2^{(2)n_1,n_2,n_3,n_4}$}\label{figa4pol7}
\end{figure}
\begin{figure}[p]
\begin{picture}(400,190)(0,-15)
\put(120,10){\rule{3pt}{140pt}}
\put(120,26){\line(-1,2){18}}
\put(120,134){\line(-1,-2){18}}
\put(120,80){\line(-1,-1){18}}
\put(120,80){\line(-1,1){18}}
\put(102,62){\line(-2,1){60}}
\put(102,98){\line(-2,-1){60}}
\put(35,65){\shortstack{\scriptsize{$2$}}}
\put(35,92){\shortstack{\scriptsize{$2$}}}
\put(105,40){\shortstack{\scriptsize{$1$}}}
\put(106,118){\shortstack{\scriptsize{$1$}}}
\put(105,70){\shortstack{\scriptsize{$4$}}}
\put(105,85){\shortstack{\scriptsize{$4$}}}
\put(125,15){\shortstack{\footnotesize{$b_{n_1,n_2,n_3,n_4}$}}}
\put(125,55){\shortstack{\footnotesize{$b_{n_1-1,n_2,n_3,n_4}$}}}
\put(125,100){\shortstack{\footnotesize{$b_{n_1-1,n_2,n_3,n_4}$}}}
\put(125,140){\shortstack{\footnotesize{$b_{n_1,n_2,n_3,n_4}$}}}
\put(125,128){\shortstack{\tiny{$\(\frac 12-(2n_1-1)\frac B2\)$}}}
\put(125,74){\shortstack{\tiny{$\(\frac 52+(2n_1-1)\frac B2\)$}}}
\put(320,10){\rule{3pt}{140pt}}
\put(320,26){\line(-1,1){18}}
\put(320,134){\line(-1,-1){18}}
\put(320,80){\line(-1,-2){18}}
\put(320,80){\line(-1,2){18}}
\put(302,44){\line(-2,-1){60}}
\put(302,116){\line(-2,1){60}}
\put(235,10){\shortstack{\scriptsize{$2$}}}
\put(235,143){\shortstack{\scriptsize{$2$}}}
\put(308,28){\shortstack{\scriptsize{$1$}}}
\put(308,127){\shortstack{\scriptsize{$1$}}}
\put(306,61){\shortstack{\scriptsize{$3$}}}
\put(306,94){\shortstack{\scriptsize{$3$}}}
\put(325,15){\shortstack{\footnotesize{$b_{n_1,n_2,n_3,n_4}$}}}
\put(325,55){\shortstack{\footnotesize{$b_{n_1-1,n_2,n_3,n_4}$}}}
\put(325,100){\shortstack{\footnotesize{$b_{n_1-1,n_2,n_3,n_4}$}}}
\put(325,140){\shortstack{\footnotesize{$b_{n_1,n_2,n_3,n_4}$}}}
\put(325,128){\shortstack{\tiny{$\(\frac 12-(2n_1-1)\frac B2\)$}}}
\put(325,74){\shortstack{\tiny{$\(\frac 32+(2n_1-1)\frac B2\)$}}}
\put(45,5){\shortstack{\scriptsize{$\left(\frac 32 - (2n_1-1)\frac B2\right)$}}}
\put(245,5){\shortstack{\scriptsize{$\left(\frac 52 + (2n_1-1)\frac B2 \right)$}}}
\put(110,-15){\shortstack{(i)}}
\put(310,-15){\shortstack{(j)}}
\end{picture}
\caption{\it Poles in $K_2^{(2)n_1,n_2,n_3,n_4}$ with $n_1\leq 0$.}\label{figa4pol8}
\end{figure}
\begin{figure}[p]
\begin{picture}(380,190)(0,-15)
\put(120,10){\rule{3pt}{140pt}}
\put(120,55){\line(-1,-2){15}}
\put(120,105){\line(-1,2){20}}
\put(95,25){\shortstack{\scriptsize{$1$}}}
\put(95,150){\shortstack{\scriptsize{$1$}}}
\put(125,30){\shortstack{\footnotesize{$b_{-n_2,n_2,n_3,n_4}$}}}
\put(125,77){\shortstack{\footnotesize{$b_{-n_2-1,n_2+1,n_3,n_4}$}}}
\put(125,125){\shortstack{\footnotesize{$b_{-n_2,n_2,n_3,n_4}$}}}
\put(280,10){\rule{3pt}{140pt}}
\put(280,110){\line(-1,-2){42}}
\put(280,50){\line(-1,2){47}}
\put(230,25){\shortstack{\scriptsize{$1$}}}
\put(230,150){\shortstack{\scriptsize{$1$}}}
\put(285,30){\shortstack{\footnotesize{$b_{-n_2,n_2,n_3,n_4}$}}}
\put(285,77){\shortstack{\footnotesize{$b_{-n_2+1,n_2-1,n_3,n_4}$}}}
\put(285,125){\shortstack{\footnotesize{$b_{-n_2,n_2,n_3,n_4}$}}}
\put(45,5){\shortstack{\scriptsize{$\left(\frac 32 - (2n_2+1)\frac B2\right)$}}}
\put(205,5){\shortstack{\scriptsize{$\left(\frac 32 - (2n_2-1)\frac B2 \right)$}}}
\put(110,-15){\shortstack{(g)}}
\put(270,-15){\shortstack{(a)}}
\end{picture}
\caption{\it Poles in $K_1^{(2)n_1,n_2,n_3,n_4}$ if $n_1=-n_2$.}\label{figa4pol9}
\end{figure}
\begin{figure}
\begin{picture}(480,190)(0,-15)
\put(120,10){\rule{3pt}{140pt}}
\put(120,26){\line(-1,1){18}}
\put(120,134){\line(-1,-1){18}}
\put(120,80){\line(-1,-2){18}}
\put(120,80){\line(-1,2){18}}
\put(102,44){\line(-2,-1){60}}
\put(102,116){\line(-2,1){60}}
\put(35,10){\shortstack{\scriptsize{$2$}}}
\put(35,145){\shortstack{\scriptsize{$2$}}}
\put(108,28){\shortstack{\scriptsize{$1$}}}
\put(108,127){\shortstack{\scriptsize{$1$}}}
\put(106,61){\shortstack{\scriptsize{$3$}}}
\put(106,94){\shortstack{\scriptsize{$3$}}}
\put(125,15){\shortstack{\footnotesize{$b_{-n_2,n_2,n_3,n_4}$}}}
\put(125,55){\shortstack{\footnotesize{$b_{-n_2+1,n_2-1,n_3,n_4}$}}}
\put(125,140){\shortstack{\footnotesize{$b_{-n_2,n_2,n_3,n_4}$}}}
\put(125,128){\shortstack{\tiny{$\(\frac 32-(2n_2-1)\frac B2\)$}}}
\put(125,74){\shortstack{\tiny{$\(\frac 12+(2n_2-1)\frac B2\)$}}}
\put(320,10){\rule{3pt}{140pt}}
\put(320,26){\line(-1,2){18}}
\put(320,134){\line(-1,-2){18}}
\put(320,80){\line(-1,-1){18}}
\put(320,80){\line(-1,1){18}}
\put(302,62){\line(-2,1){60}}
\put(302,98){\line(-2,-1){60}}
\put(235,65){\shortstack{\scriptsize{$2$}}}
\put(235,92){\shortstack{\scriptsize{$2$}}}
\put(305,40){\shortstack{\scriptsize{$1$}}}
\put(306,118){\shortstack{\scriptsize{$1$}}}
\put(305,70){\shortstack{\scriptsize{$4$}}}
\put(305,85){\shortstack{\scriptsize{$4$}}}
\put(325,15){\shortstack{\footnotesize{$b_{-n_2,n_2,n_3,n_4}$}}}
\put(325,55){\shortstack{\footnotesize{$b_{-n_2+1,n_2-1,n_3,n_4}$}}}
\put(325,140){\shortstack{\footnotesize{$b_{n_1,n_2,n_3,n_4}$}}}
\put(325,128){\shortstack{\tiny{$\(\frac 32-(2n_2-1)\frac B2\)$}}}
\put(325,74){\shortstack{\tiny{$\(\frac 32+(2n_2-1)\frac B2\)$}}}
\put(45,5){\shortstack{\scriptsize{$\left(\frac 32 + (2n_2-1)\frac B2\right)$}}}
\put(245,5){\shortstack{\scriptsize{$\left(\frac 52 - (2n_2-1)\frac B2 \right)$}}}
\put(110,-15){\shortstack{(d)}}
\put(310,-15){\shortstack{(e)}}
\end{picture}
\caption{\it Poles in $K_2^{(2)n_1,n_2,n_3,n_4}$ if $n_1=-n_2$.}\label{figa4pol10}
\end{figure}
Again the reflection amplitudes for particles $3$ and $4$ are
obtained by charge conjugation. Notice that these reflection
amplitudes for the second boundary condition contain those for the
first boundary condition as a factor. We only need to explain the
poles introduced by the new additional factors. These all depend on
$n_1$ or $n_4$ and must thus be due to diagrams involving the
emission or absorption of particles $1$ or $4$ from the boundary.
For $K_1^{(2)n_1,n_2,n_3,n_4}(\th)$ the relevant diagrams are
displayed in figure~\ref{figa4pol6}, while for
$K_2^{(2)n_1,n_2,n_3,n_4}(\th)$ they are given in figure~\ref{figa4pol7}.
Diagram \ref{figa4pol7}(j) exists only if $n_4>0$. Otherwise the
corresponding pole lies off the physical sheet.
If $n_1\leq0$ then diagram \ref{figa4pol7}(i) changes to diagram
\ref{figa4pol8}(i). Also if $n_1\leq 0$ then the second pole in line
(j) is on the physical sheet, and corresponds to diagram
\ref{figa4pol8}(j).
It is a fun exercise to go through all the special
cases for which blocks in the reflection amplitudes coincide. For
example if $n_4+n_1=0$ then line (i) cancels
against line (l). Correspondingly figures \ref{figa4pol7}(i) and
\ref{figa4pol8}(i) no longer
provide poles because the reflection amplitudes in the middle of
those diagrams have double zeros. The disappearance of the zero in
line (l) at $n_4+n_1=0$ in turn has the effect that diagrams
\ref{figa4pol7}(j) and \ref{figa4pol8}(j) provide double poles if
$n_4+n_1=1$ which is correct because then the two blocks on line (j)
coincide.
If $n_1+n_2=0$ then the pole on line (f) in $K_1^{(2)n_1,n_2,n_3,n_4}$
disappears and thus diagram \ref{figa4pol6}(f) also has to
disappear. That means that we need to exclude any boundary bound
states with
$n_1+n_2<0$. From the charge conjugated amplitudes we see that
we also have to exclude states with $n_4+n_3<0$. This is the
justification for the restriction on the lattice of states
mentioned above.
The poles in line (g) and (a) in $K_1^{(2)n_1,n_2,n_3,n_4}$ have a new
explanation if $n_1+n_2=0$ in terms of the diagrams in figure
\ref{figa4pol9}, and diagram \ref{figa4pol1}(c) and its corresponding
pole disappear.
At $n_1+n_2=0$ also the diagrams \ref{figa4pol2}(b), \ref{figa4pol3}(d),
\ref{figa4pol4}(d), \ref{figa4pol3}(e), \ref{figa4pol7}(i)
\ref{figa4pol8}(i) and \ref{figa4pol8}(j) disappear. Most of the
corresponding poles also
disappear but the first poles on lines (d) and (e) do not and are
explained by the two new diagrams in figure~\ref{figa4pol10}.
We should still draw more diagrams to explain new poles which
become relevant if one of the indices becomes sufficiently
negative. However we imagine that the reader will be exhausted by
now and so are we and we will therefore stop here. Clearly what is
needed is a more systematic treatment which could explain why
everything falls into place so miraculously.
\section{Discussion}
The most striking result of this paper is probably the fact that the
boundary of real coupling affine Toda theory has such a rich spectrum
of boundary states. Already in the simplest case of $a_2^{(1)}$
Toda theory the spectrum is rather intricate, as displayed in
figures \ref{spectrum1} and \ref{spectrum2}. The pole structure of
the reflection amplitudes necessitates the existence of these
states. If one of them were absent there would be unexplained
poles. However we can not exclude the possibility that there might
be even more states. There are many simple poles in the reflection
amplitudes which we were able to explain by anomalous threshold
diagrams and for which we therefore did not have to introduce
new boundary states. However it is possible
that some of the anomalous threshold poles mask genuine bound
state poles. This does happen in the S-matrices of non self-dual
affine Toda theories \cite{deliu92,corri93}.
Whether it happens in $a_n^{(1)}$ Toda theory with a boundary
could only be decided if one were
able to calculate the residues of the anomalous threshold poles
and compare them to the actual residues of the poles in the
reflection amplitudes to see if there is a mismatch. Such calculations
were carried out in perturbation theory for the Toda S-matrices
\cite{brade91,deliu92} but we do not know how to perform such calculations
in the presence of the boundary.
Another result of this paper, which was not obvious from the
classical analysis, is
that many classically different integrable boundary conditions are
equivalent at the quantum level, in the sense that they give rise
to the same boundary states and the same reflection amplitudes.
When the classical real-valued vacuum solutions of
$a_n^{(1)}$ Toda theory with solitonic boundary conditions
were obtained in
\cite{deliu98} it was remarked that they contain a
continuous parameter. We were wondering what consequence this
vacuum degeneracy would have in the quantum theory. For the
$(- -\cdots- -)$ boundary condition of $a_{2n+1}^{(1)}$ Toda
theory
this degeneracy had been observed earlier in \cite{fujii95}.
In that paper it was implied that the vacuum degeneracy implies
that the theory is ``unstable''. We disagree with that
interpretation. The present paper shows that it is consistent
to assume that the quantum theory picks a unique stable vacuum state.
We have derived reflection amplitudes for those integrable
boundary conditions which satisfy the constraint that the product
of all boundary parameters is $+1$. These are the boundary
conditions for which non-singular real vacuum solutions can be
obtained by analytic continuation of stationary soliton solutions.
We have no results for those boundary conditions for which the
product of all boundary parameters is $-1$.
The Durham group had made a conjecture in \cite{corri94} for the
boundary conditions which has all $C_i=-1$. For $n$ odd this is
a solitonic boundary condition of type $(n+1)/2$
and we can compare their conjecture
with our result. Not surprisingly our results differ because the
calculations in \cite{corri94} were based on a constant vacuum,
whereas we use a vacuum obtained by continuing a stationary soliton
solution, which has the same energy as the constant solution.
So these theories appear to have two different vacuum sectors.
For the $(++\cdots++)$ boundary condition on the other hand the
constant solution is the lowest energy solution and
the conjecture made in \cite{corri94} does agree with
our result \eqref{Ka}.
The results of this paper are of course not rigorous. At several
stages we had to make assumptions. We will list the most important
ones here so that the reader can decide how much confidence he
wants to put into them.
1) We have assumed that the scalar factor
of the soliton reflection matrix is given by \eq{aa} which was the
minimal solution of the equations \eqref{aeq}. However there is
the freedom of including CDD factors of the form \eqref{cdd}. We
believe that inclusion of any of these CDD factors would lead to
an inconsistent pole structure after bootstrap, but we have not
performed a careful analysis of this. The semiclassical
comparision between or reflection matrix and the classical time
delay which we have performed does not detect the CDD factors.
2) We have assumed that the classical solutions to the solitonic boundary
conditions found in \cite{deliu98} are indeed the solutions of
lowest energy. The calculations performed during the work on
\cite{deliu98} have given us strong confidence that this is indeed
the case but we don't know of any way to prove it.
3) The classical vacuum solutions which we use are obtained by
analytic continuation to real coupling of solutions describing a
stationary soliton in front of the boundary in the imaginary
coupling theory. We have assumed that therefore also the particle
reflection amplitudes in the real coupling theory are obtained by
the same analytic continuation of the corresponding breather
reflection amplitude in the imaginary coupling theory. This
assumption can really only be justified a posteriori by the fact
that it has led to consistent results.
Clearly, we have not exhausted the subject of Toda reflection
amplitudes in this paper. While we have indeed given the
reflection amplitudes for all known vacuum solutions satisfying
integrable boundary conditions, we have completely worked out the
reflection
amplitudes for the excited boundary states only for two theories.
We have not discussed the special coincidences and
truncations which can occur at special rational values of the
coupling constant. We have given an example of such a phenomenon
at the end of section \ref{subsect:mixed2} where a coupling constant
$b=\frac{1}{2n}$ with $n$ any integer leads to a degeneracy in the
spectrum.
Furthermore, we have restricted our attention to the regime where
the coupling constant satisfies $B<1$. The spectrum of bound
states changes drastically as soon as $B$ equals $1$.
The study of the case $B>1$ is related to the question of
weak--strong duality. Affine Toda theory in the bulk displays a
fascinating weak--strong coupling duality under which
$\beta\leftrightarrow\frac{4\pi}{\beta}$ or $B\leftrightarrow
2-B$. In particular the S-matrices of $a_n^{(1)}$ Toda theory
are self-dual under
this transformation of the coupling constant.
Now that we have the reflection amplitudes we can study the
interesting question of how S-duality works in the presence of
boundaries. The only thing which we can say so far is that the
Neumann boundary condition is dual to the $(++\cdots++)$ boundary
condition. Previously this duality had been observed in the
sine-Gordon model \cite{corri97} and the
$a_2^{(1)}$ theory \cite{gande98b}.
Without having to do any detailed comparisons we can immediately
say that none of the reflection amplitudes which we have
calculated coincide with the reflection amplitudes which had been
conjectured in \cite{fring94,kim95b,kim,sasak93}. These papers had assumed that the
reflection amplitudes should be invariant under $B\rightarrow 2-B$
which ours manifestly are not.
Working out the explanation of the pole structure of the
reflection amplitudes in the examples of $a_2^{(1)}$ and
$a_4^{(1)}$ Toda theory has been rather tedious.
However it has been quite fascinating to see the
miraculous way in which
these Toda theories have managed to be consistent.
But miracles demand to be studied and (eventually) understood.
This probably will require more cases for more different algebras
to be worked out. It should be possible to give explicit formulas
for the particle reflection factors off the excited boundary states
in $a_n^{(1)}$ theories for
general $n$, starting from our formula \eqref{Kmixed}
for the particle
reflection amplitudes off the vacuum boundary state.
To extend the results to algebras other than $a_n^{(1)}$ one will
have to repeat the analysis of section \ref{sect:an} to find the
soliton reflection matrices as solutions of the corresponding
reflection equations
and then the breather reflection amplitudes
by bootstrap. To formulate the reflection equations one will need
to use the soliton S-matrices which are known for many algebras
\cite{gande96}. We suspect that again the simplest diagonal
solutions of the reflection equations will correspond to the
uniform $(++\cdots ++)$
boundary condition.
Because in all studied cases it has been found that
the particle S-matrices are obtained from the breather S-matrices
by analytical continuation to real coupling, it is to be expected
that the same will hold for the reflection
amplitudes.
To find the reflection amplitudes for other integrable boundary
conditions we used our knowledge of the corresponding vacuum
solutions from \cite{deliu98}. The results of that paper can also
be used to find vacuum solutions for the solitonic boundary conditions
in $c_n^{(1)}$ Toda theories
and therefore the analysis of this paper could be carried over to
those cases.
Clearly one would like to eventually go beyond a case-by-case
analysis and obtain general formulas for the reflection
amplitudes. Would formulas as elegant as those provided by Dorey
\cite{dor} for the S-matrices in terms of root systems be possible?
We hope that the exact results for the reflection amplitudes,
which we have given in this paper, will be a useful
guide in the attempts to generalise standard quantum field theoretic
techniques to field theories with boundaries.
For example, perturbation theory in the presence of a non-constant
boundary background is a non-trivial matter.
So far only one-loop calculations for constant backgrounds
have been performed \cite{kim,perki98} and our
results confirm them.
\vspace{1cm}\noindent{\bf\large{Acknowledgements}}
We would like to thank Peter Bowcock, Edward Corrigan, Patrick Dorey
and Gerard Watts for interesting discussions.
GMG has been supported by EPSRC research grant no. GR/K 79437 and
GWD by an EPSRC advanced fellowship. This collaboration has been
supported in part by EU contract FMRX-CT96-0012.
\parskip 1pt
{\footnotesize
|
\section{Introduction}
We report on an experimental study of hadronic shower profiles detected by
the prototype of the ATLAS Barrel Tile Hadron Calorimeter (Tile calorimeter)
\cite{atcol94}, \cite{tilecal-tdr96}.
The innovative design of this calorimeter, using longitudinal segmentation
of active and passive layers (see Fig.\ \ref{fig:f1}), provides an
interesting system for the measurement of hadronic shower profiles.
Specifically, we have studied the transverse development of hadronic showers
using 100 GeV pion beams and longitudinal development of hadronic showers
using 20 -- 300 GeV pion beams.
Characteristics of shower development in hadron calorimeters have been
published for some time.
However, a complete quantitative description of transverse
and longitudinal properties of hadronic showers does not exist
\cite{bock97}.
The transverse profiles are usually expressed as a
function of transverse coordinates, not the radius,
and are integrated over the other coordinate
\cite{womersley88}.
The three-dimensional parametrization of hadronic showers described here
could be a useful starting point for fast simulations,
which can be faster than full simulations at the microscopic level
by several orders of magnitude
\cite{bock81}, \cite{grindhammer90}, \cite{brun91}.
The paper is organised as follows.
In Section 2, the calorimeter and the test beam setup are briefly described.
In Section 3, the mathematical procedures for extracting the underling
radial energy density of hadronic showers are developed.
The obtained results on the transverse and longitudinal profiles,
the radial energy densities and the radial containment
of hadronic shower are presented in Sections 4 -- 7.
Section 8 and 9 investigate the three-dimensional parametrization
and electromagnetic fraction of hadronic shower.
Finally Section 10 contains a summary and the conclusions.
\section{The Calorimeter}
The prototype Tile Calorimeter used for this study is composed
of five modules
stacked in the $Y$ direction, as shown in Fig.\ \ref{fig:f01}.
Each module spans $2 \pi / 64$ in the azimuthal angle,
100 cm in the $Z$ direction,
180 cm in the $X$ direction (about 9 interaction lengths,
$\lambda_I$, or about 80
effective radiation lengths, $X_{0}$), and has a front face of 100
$\times$ 20 cm$^2$
\cite{berger95}.
The absorber structure of each module consists of 57 repeated ``periods".
Each period is 18 mm thick and consists of four layers.
The first and third layers are formed by large trapezoidal steel plates
(master plates) and span the full longitudinal dimension of the module.
In the second and fourth layers, smaller trapezoidal steel plates
(spacer plates) and scintillator tiles alternate along the $X$ direction.
These layers consist of 18 different trapezoids of steel or scintillator,
each of 100 mm in depth.
The master plates, spacer plates and scintillator tiles are
5 mm, 4 mm and 3 mm thick, respectively.
The iron to scintillator ratio is $4.67 : 1$ by volume.
The calorimeter thickness along the beam direction at the incidence angle
(the angle between the incident particle direction and the normal to the
calorimeter front face) of $\Theta = 10^{\circ}$ corresponds to 1.49 m
of iron equivalent
\cite{lokajicek95-64}.
Wavelength shifting fibers collect scintillation light from the
tiles at both of their open (azimuthal) edges and bring it to
photo-multipliers
(PMTs) at the periphery of the calorimeter.
Each PMT views a specific group of tiles through the corresponding
bundle of fibers.
The modules are divided into five segments along $Z$.
They are also longitudinally segmented (along $X$) into four depth segments.
The readout cells have a lateral dimensions of 200 mm along $Z$,
and longitudinal
dimensions of 300, 400, 500, 600 mm for depth segments 1 -- 4, corresponding
to 1.5, 2,
2.5 and 3 $\lambda_{I}$ at $\Theta = 0^{\circ}$ respectively.
Along $Y$, the cell sizes vary between about 200 and 370 mm depending on
the X (depth) coordinate.
We recorded energies for 100 different cells for each event
\cite{berger95}.
The calorimeter was placed on a scanning table that allowed
movement in any direction.
Upstream of the calorimeter, a trigger counter telescope
(S1 -- S3) was installed,
defining a beam spot approximately 20 mm in diameter.
Two delay-line wire chambers (BC1 -- BC2),
each with $(Z, Y)$ readout,
allowed the impact point of beam particles on the
calorimeter face to be reconstructed to better than $\pm 1$ mm
\cite{ariztizabal94}.
``Muon walls'' were placed behind (800$\times$800 mm$^2$)
and on the positive $Z$ side (400$\times$1150 mm$^2$) of the calorimeter
modules
to measure longitudinal and lateral hadronic shower leakage
\cite{lokajicek95-63}.
We used the TILEMON program \cite{efthymiopoulos95} to convert the
raw calorimeter data into PAW Ntuples \cite{paw} containing
calibrated cell energies and other information used in this study.
The data used for the study of lateral profiles were collected
in 1995 during
a special $Z$-scan run at the CERN SPS test beam.
The calorimeter was exposed to 100 GeV negative pions at a $10^{\circ}$
angle with varying impact points in the $Z$-range from $- 360$ to $+ 200$ mm.
A total of $>$~300,000 events have been analysed;
for the lateral profile study only events without lateral leakage were
used.
The uniformity of the calorimeter's response for this $Z$-scan is
estimated
to be $1\%$
\cite{budagov96-76}.
The data used for the study of longitudinal profiles were obtained using
20 -- 300 GeV negative pions at a $20^{\circ}$ angle and were also taken
in 1995
during the same test beam run.
\section{Extracting the Underlying Radial Energy Density}
In this investigation we use a coordinate system based on the incident
particle direction.
The impact point of the incident particle at the calorimeter front face
defines the origin of coordinate system.
The incident particle direction forms the $x$ axis,
while the $y$ axis is in the same direction as $Y$ defined in Section 2.
The normal to the $xy$ surface
defines
the $z$ axis.
We measure the energy deposition in each calorimeter cell for every event.
In the $ijk$-cell of the calorimeter with the volume $V_{ijk}$ and cell
center coordinates $(x_{c},y_{c},z_{c})$, the energy deposition $E_{ijk}$ is
\begin{equation}
E_{ijk}(x_{c},y_{c},z_{c}) =
\int \limits_{V_{ijk}} f(x,y,z)\ dx dy dz,
\label{e010}
\end{equation}
where $f(x,y,z)$ is the three-dimensional hadronic shower
energy density function.
Due to the azimuthal symmetry of shower profiles, the density $f(x,y,z)$ is
only a function of the radius $r = \sqrt{y^2+z^2}$ from the shower axis
and the longitudinal coordinate $x$.
Then
\begin{equation}
E_{ijk}(x_{c},y_{c},z_{c}) =
\int \limits_{V_{ijk}} \Psi(x, r)\ rdr d \phi dx,
\label{e010-01}
\end{equation}
where $\phi$ is the azimuthal angle and
$\Psi(x,r)$ has the form of a joint probability density function
(p.d.f.)
\cite{review96}.
The joint p.d.f. can be further decomposed as a product of
the marginal p.d.f., $dE (x)/ dx$,
and the conditional p.d.f., $\Phi (x, r)$,
\begin{equation}
\Psi(x,r) =
\frac{dE(x)}{dx} \cdot \Phi (x,r).
\label{e123-10}
\end{equation}
The longitudinal density $dE/dx$ is defined as
\begin{equation}
\frac{dE(x)}{dx} =
\int \limits_{-\infty}^{\infty}
\int \limits_{-\infty}^{\infty}
f(x,y,z)\ dy dz .
\label{e08}
\end{equation}
Finally, the radial density function $\Phi(r)$ for a given depth segment is
\begin{equation}
\Phi(r) = E_{0}\ \Phi(x,r) ,
\label{e123-001}
\end{equation}
where $E_{0}$ is the total shower energy deposition into the depth segment
for fixed $x$.
There are several methods for extracting the radial density $\Phi(r)$
from the measured distributions of energy depositions.
One method is to unfold $\Phi(r)$ using expression (\ref{e010-01}).
This method was used in the analysis of the data from the lead-scintillating
fiber calorimeter \cite{acosta92}.
Several analytic forms of $\Phi(r)$ were tried,
but the simplest that describes the
energy deposition in cells was a combination of an exponential
and a Gaussian:
\begin{equation}
\Phi(r) =
\frac{b_1}{r}\ e^{ - \frac{r}{\mu_1}} +
\frac{b_2}{r}\ e^{ - {( \frac{r}{\mu_{2}} )}^2 },
\label{e1}
\end{equation}
where $b_i$ and $\mu_i$ are the free parameters.
Another method for extracting the radial density
is to use the marginal density function
\begin{equation}
f(z) =
\int \limits_{-\infty}^{\infty}
\int \limits_{x_1}^{x_2}
f(x,y,z)\ dx dy .
\label{e09-1}
\end{equation}
which is related to the radial density $\Phi(r)$
\begin{equation}
f(z) =
2 \int \limits_{|z|}^{\infty}
\frac{\Phi(r)\ r dr}{\sqrt{r^{2}-z^{2}}} .
\label{e4}
\end{equation}
This method was used \cite{lednev95}
for extracting the electron shower transverse profile
from the GAMS-2000 electromagnetic calorimeter data
\cite{akopdijanov77}.
The above integral equation (\ref{e4}) can be reduced to an Abelian equation
by replacing variables
\cite{whitteker27}.
In \cite{budagov97},
the following solution to equation (\ref{e4}) was obtained
\begin{equation}
\Phi(r) =
- \frac{1}{\pi}\ \frac{d}{dr^2}
\int \limits_{r^2}^{\infty}
\frac{f(z)\ d z^2}{\sqrt{z^{2}-r^{2}}}.
\label{e5}
\end{equation}
For our study, we used the sum of three exponential functions to
parameterize $f(z)$ as
\begin{equation}
f(z) =
\frac{E_{0}}{2B}\ \sum_{i=1}^{3}\ a_i\
e^{ - \frac{|z|}{{\lambda}_{i}} },
\label{e21}
\end{equation}
where
$z$ is the transverse coordinate,
$E_{0},\ a_i,\ \lambda_i$ are free parameters,
$B = \sum_{i=1}^{3} a_i \lambda_i$,
$\sum_{i=1}^{3} a_i = 1$
and $\int_{-\infty}^{+\infty} f (z) dz = E_{0}$.
In this case, the radial density function, obtained by integration
and differentiation of equation (\ref{e5}), is
\begin{equation}
\Phi(r) =
\frac{E_{0}}{2 \pi B}\
\sum_{i=1}^{3}\ \frac{a_{i}}{\lambda_{i}}\
K_{0} \biggl( \frac{r}{\lambda_{i}} \biggr),
\label{e23}
\end{equation}
where
$K_{0}$ is the modified Bessel function.
This function goes to $\infty$\ as $r \rightarrow 0$\
and goes to zero as $r \rightarrow \infty$.
We define a column of five cells in a depth segment as a tower.
Using the parametrization shown in equation (\ref{e21}),
we can show that the energy deposition in a tower \cite{gavrish91},
$E(z) = \int_{z - h/2}^{z + h/2} f(z) dz$,
can be written as
\begin{eqnarray}
E(z) =
& E_{0} -
\frac{E_{0}}{B}\ \sum \limits_{i=1}^{3}\ a_i {\lambda}_i\
\cosh (\frac{ |z| }{ {\lambda}_i }) \
e^{ - \frac{h}{2 {\lambda}_i} },
& for\ |z| \leq \frac{h}{2},
\\
E(z) =
& \frac{E_{0}}{B}\ \sum \limits_{i = 1}^{3}\ a_i {\lambda}_i\
\sinh (\frac{h}{2 \lambda_i }) \
e^{ - \frac{|z|}{{\lambda}_i}},
& for\ |z| > \frac{h}{2},
\label{e023}
\end{eqnarray}
where $h$ is the size of the front face of the tower along the $z$ axis.
Note that as $h \rightarrow 0$, we get $E (z) / h \rightarrow f(z)$.
As $h \rightarrow \infty $, we find that $E (0) \rightarrow E_{0}$.
The full width at half maximum (FWHM)
of an energy deposition profile for small values of $(FWHM - h)/h$
can be approximated by
\begin{equation}
FWHM = h +
\frac{2E(h/2) - E(0)}{
\frac{E_{0}}{B}\
\sum \limits_{i=1}^{3}\ a_i\
\sinh ( \frac{h}{2 \lambda_i }) \
e^{- \frac{h}{2 \lambda_i}}
}.
\label{e0023}
\end{equation}
We will show below that this approximation agrees well with our data.
A cumulative function may be derived from the density function as
\begin{equation}
F(z) =
\int \limits_{-\infty}^{z} f(z)\ dz .
\label{e3-01}
\end{equation}
For our parametrization in equation (\ref{e21}), the cumulative
function becomes
\begin{eqnarray}
F(z) =
& \frac{E_{0}}{2B}\ \sum \limits_{i=1}^{3}\ a_i\ \lambda_i\
e^{ \frac{z}{ \lambda_{i} } },
& for\ z \leq 0,
\\
F(z) =
& E_{0} -
\frac{E_{0}}{2B}\ \sum \limits_{i=1}^{3}\ a_i\ \lambda_i\
e^{- \frac{z}{\lambda_{i}}},
& for\ z > 0,
\label{e23-2}
\end{eqnarray}
where
$z$ is the position of the edge of a tower along the $z$ axis.
Note that the cumulative function does not depend on the cell size
$h$.
We can construct the cumulative function and deconvolute
the density f(z) from it for any size calorimeter cell.
Note also that cumulative function is well behaved at the key points:
$F( - \infty ) = 0$,\ $F( 0 ) = E_{0} / 2$,\
and $F( \infty ) = E_{0}$.
The radial containment of a shower as a function of $r$ is given by
\begin{equation}
I(r) =
\int \limits_0^r \int \limits_0^{2 \pi}
\Phi (r)\ r dr d\phi =
E_{0} -
\frac{E_{0} r}{B}\ \sum \limits_{i=1}^{3}\ a_{i}\
K_1 \biggl( \frac{r}{{\lambda}_{i}} \biggr),
\label{e008}
\end{equation}
where $K_{1}$ is the modified Bessel function.
As $r \rightarrow \infty$, the function
$r K_1 (r)$ tends to zero and we get $I(\infty) = E_{0}$, as expected.
We use two methods to extract the radial density function
$\Phi(r)$.
One method is to unfold $\Phi(r)$ from (\ref{e010-01}).
Another method is to use the expression (\ref{e5}) after
we have obtained the marginal density function $f(z)$.
There are three ways to extract $f(z)$:
by fitting the energy deposition $E(z)$ \cite{gavrish91},
by fitting the cumulative function $F(z)$ and
by directly extracting $f(z)$ by numerical differentiation of
the cumulative function.
The effectiveness of these various methods depend on the
scope and quality of the experimental data.
\section{Transverse Behaviour of Hadronic Showers}
Figure \ref{fig:f5-001} shows the energy depositions in towers
for depth segments 1 -- 4 as a function of the $z$ coordinate of the
center of the tower.
Figure \ref{fig:f6} shows the same for the entire calorimeter
(the sum of the histograms presented in
Fig.\ \ref{fig:f5-001}).
These Figures are profile histograms \cite{paw} and
give the energy deposited in any tower
for all analysed events in bins of the $z$ coordinate.
Here the coordinate system is linked to the incident particle direction
where $z=0$ is the coordinate of the beam impact points at the
calorimeter front face.
Figure \ref{fig:f01-1}
schematically shows a top view of the experimental setup and indicates the
$z$ coordinate of each tower.
The $z_1$, $z_2$, $z_3$ and $z_4$
are the distances between the centre of towers
(for the four depth segments) and the direction of beam
particle.
The values of the $z$-coordinate of the tower centers are negative
to the left of the beam, positive to the right of the beam and
range from $- 750$ mm to $+ 600$ mm.
To avoid edge effects, we present tower energy depositions in the range
from $- 650$ mm to $+ 500$ mm.
Note that the total tower height
(about 1.0 m at the front face, and about 1.8 m at the back)
is sufficient for shower measurements without significant leakage
in the vertical direction.
As mentioned earlier, events with significant lateral leakage
(identified by a clear minimum-ionising signal
in the lateral muon wall) were discarded. The resulting left-right
asymmetries in the distributions of Figures \ref{fig:f5-001}
and \ref{fig:f6} are very small.
As will be shown later (Section 6), the $99\%$ containment
radius is less than $500$ mm.
The fine-grained $z$-scan provided many different beam
impact locations within the ca\-lorimeter.
Due to this, we obtained a detailed picture of the
transverse shower behaviour in the calorimeter.
The tower energy depositions shown in
Figures \ref{fig:f5-001} and \ref{fig:f6}
span a range of about three orders of magnitude.
The plateau for $|z| < 100$ mm ($h/2$) and the fall-off at large $|z|$
are apparent.
Similar behaviour of the transverse profiles was observed in other
calorimeters as well
\cite{acosta92}, \cite{gavrish91}.
We used the distributions in Figs.\ \ref{fig:f5-001} and \ref{fig:f6}
to extract the underlying
mar\-gi\-nal densities function for four depth segments of the calorimeter
and for the entire calorimeter.
The solid curves in these figures
are the results of the fit with equations (12) and (\ref{e023}).
The fits typically differ from the experimental distribution by less
than $5\%$.
In comparison with
\cite{gavrish91} and \cite{binon83},
where the transverse profiles exist only for distances less than
$250$ mm, our more extended profiles (up to $650$ mm) require that
the third exponential term be introduced.
The parameters $a_i$ and $\lambda_i$, obtained by fitting, are listed in
Table~\ref{Tb2}.
The values of the parameter $E_{0}$, the average energy shower deposition
in a given depth segment,
are listed in Table~\ref{long-1}.
We have compared our values of $\lambda_1$ and $\lambda_2$
with the ones from the conventional iron-scintillator calorimeter
described in \cite{binon83}.
At $100$ GeV, our results for the entire calorimeter are
$23\pm1$ mm and $58\pm4$ mm for $\lambda_1$ and $\lambda_2$ respectively.
They agree well with the ones from \cite{binon83},
which are $18\pm3$ mm and $57\pm4$ mm.
We determined the FWHM of energy deposition profiles
(Figs.\ \ref{fig:f5-001} and \ref{fig:f6}) using formula (\ref{e0023}).
The characteristic FWHM are found to be approximately equal to
transverse tower size.
The relative difference of FWHM from
transverse tower size, $(FWHM - h)/h$, amount to $2\%$
for depth segment 1, depth segment 2 and for the entire calorimeter,
$7\%$ for depth segment 3 and $15\%$ for depth segment 4.
Figure \ref{fig:5-00} shows the calculated marginal density
function $f(z)$ and the energy deposition function, $E(z)/h$, at various
transverse sizes of tower $h = 50,\ 200,\ 300$ and 800 mm
using the obtained parameters for the entire calorimeter.
As a result of the volume integration,
the sharp $f(z)$ is transformed to the wide function $E(z)/h$,
which clearly shows its relationship to the transverse width of a tower.
The values of FWHM are 40 mm for $f(z)$ and 204 mm for $E(z)/h$
at $h = 200$ mm.
Note that the transverse dimensions of a tower vary from 300 mm to 800 mm
for the different depth segments in final ATLAS Tile calorimeter design.
The difference between $f(0)$ and $E(0)/h$ becomes less then $5\%$ only
at $h$ less then 6 mm.
The parameters $a_i$ and $\lambda_i$
as a function of $x$ (in units of $\lambda_{\pi}^{Fe}$)
are displayed in Figs.\ \ref{fig:5} and \ref{fig:05}.
Here $\lambda_{\pi}^{Fe} = 207$ mm is the nuclear interaction length
for pions in iron
\cite{budagov97}.
In these calculations, the effect of the $10^{\circ}$ incidence beam
angle has been corrected.
As can be seen from Figs.\ \ref{fig:5} and \ref{fig:05},
the value of $a_1$ decreases and the values of the remaining parameters,
$a_2, a_3$ and $\lambda_i$, increase as the shower develops.
This is a reflection of the fact that as the hadronic shower propagates
into the calorimeter it becomes broader.
Note also that the $a_i$ and $\lambda_i$ parameters demonstrate
linear behaviour as a function of $x$.
The lines shown are fits to the linear equations
\begin{equation}
a_i (x) = \alpha_i + \beta_i x
\label{eai}
\end{equation}
and
\begin{equation}
\lambda_i (x) = \gamma_i + \delta_i x.
\label{eli}
\end{equation}
The values of the parameters
$\alpha_i$, $\beta_i$, $\gamma_i$ and $\delta_i$
are presented in Table \ref{Tb02}.
It is interesting to note that
the linear behaviour of the slope exponential was also observed
for the low-density fine-grained flash chamber calorimeter
\cite{womersley88}.
However, some non-linear behaviour of the slope of the halo component
was demonstrated for the uranium-scintillator $ZEUS$ calorimeter
at interaction lengths more than 5 $\lambda_I$
\cite{barreiro90}.
Similar results were obtained for the cumulative function distributions.
The cumulative function $F(z)$ is given by
\begin{equation}
\label{eq-pr-2}
F (z) = \sum_{k=1}^{4} F^{k}(z),
\end{equation}
where $F^{k}(z)$ is the cumulative function for depth segment $k$.
For each event, $F^{k}(z)$ is
\begin{equation}
\label{eq-pr-3}
F^{k} (z) = \sum_{i=1}^{i_{max}}
\sum_{j=1}^{5} E_{ijk},
\end{equation}
where $i_{max} = 1, \ldots, 5$ is the last tower number in the sum.
Figures \ref{fig:f7-001} and \ref{fig:f9-1} present the cumulative
functions for four depth segments and for the entire calorimeter.
The curves are fits of equations (16) and (\ref{e23-2}) to the data.
Systematic and statistical errors are again added in quadrature.
The results of the cumulative function fits are less reliable
and in what follows we use the results from energy depositions in a tower.
However, the marginal density functions determined by the two methods
(by using the energy deposition spectrum and the cumulative function)
are in reasonable agreement.
\section{Radial Hadronic Shower Energy Density}
Using formula (\ref{e23}) and the values of the parameters
$a_i$, $\lambda_i$, given in Table~\ref{Tb2},
we have determined the underlying radial
hadronic shower energy density functions, $\Phi(r)$.
The results are shown in Figure \ref{fig:f17-001} for depth segments 1 -- 4
and in Figure \ref{fig:f10-1} for the entire calorimeter.
The contributions of the three terms of $\Phi(r)$ are also shown.
The functions $\Phi(r)$ for separate depth segments of a calorimeter are given
in this paper for the first time. The function $\Phi(r)$ for the entire
calorimeter was previously given for a Lead-scintillating fiber calorimeter
\cite{acosta92}; it is given here for the first time for an Iron-scintillator
calorimeter. The analytical functions giving the radial energy density
for different depth segments allow to easily obtain
the shower energy deposition in any calorimeter cell,
shower containment fractions and the cylinder radii for any
given shower containment fraction.
The function $\Phi(r)$ for the entire calorimeter has been compared
with the one for the lead-scintillating fiber calorimeter
of ref.\ \cite{acosta92}, that has
about the same effective nuclear interaction length for pions
(namely 251 mm for the tile and 244 mm for the fiber
calorimeter \cite{budagov97}).
The two
radial density functions are rather similar as seen
in Fig.\ \ref{fig:f10-2}.
The lead-scintillating fiber calorimeter density function
$\Phi(r)$, which
was obtained from
a 80 GeV $\pi^{-}$ grid scan at an angle
of $2^{\circ}$ with respect to the fiber direction,
was parametrized
using formula (\ref{e1}) with
$b_1 = 0.169$ pC/mm, $b_2 = 0.677$ pC/mm, ${\mu}_1 = 140$ mm and
${\mu}_2 = 42.4$ mm.
For the sake of comparing the radial density functions of the two calorimeters,
the distribution from \cite{acosta92} was normalised to the $\Phi(r)$ of the
Tile calorimeter.
Precise agreement between these functions
should not be expected because of
the effect of the
different absorber materials
used in the two detectors
(e.\ g.\ the radiation/interaction length ratio for the Tile calorimeter
is three times larger than for lead-scintillating fiber calorimeter
\cite{budagov97}),
the values of $e/h$ are different, as is hadronic activity of showers because
fewer neutrons are produced
in iron than in lead \cite{Fabjan82}, \cite{gabriel94}).
\section{Radial Containment}
An other issue on which new results are presented here
is the longitudinal development of
shower transverse dimensions.
The parametrization of the radial density function, $\Phi(r)$,
was integrated to yield the shower containment as a function
of the radius, $I(r)$.
Figure \ref{fig:f11-1} shows the transverse containment of the pion shower,
$I(r)$, as a function of $r$ for four depth segments
and for the entire calorimeter.
In Table \ref{Tb3} and Fig.\ \ref{fig:f11-2}
the radii of cylinders for the given shower containment
($90\%$, $95\%$ and $99\%$) extracted from Fig.\ \ref{fig:f11-1}
as a function of depth are shown.
The centers of depth segments,
$x$, are given in units of $\lambda_{\pi}^{Fe}$.
Solid lines are the linear fits to the data:
$r(90\%) = (85 \pm 6) + (37 \pm 3)x$,
$r(95\%) = (134 \pm 9) + (45 \pm 3)x$,
$r(99\%) = (349 \pm 7) + (22 \pm 2)x$ (mm).
As can be seen, these containment radii increase linearly with depth.
Such a linear increase of $95\%$ lateral shower containment
with depth is also observed
in an other iron-scintillator calorimeter at 50 and 140 GeV
\cite{holder78}.
It is interesting to note that the shower radius for $95\%$ radial
containment for the entire calorimeter is equal to
$\lambda_{\pi}^{eff} = 251$ mm \cite{budagov97}
which justifies the frequently encountered statement that
$r(95\%) \approx \lambda_I$ \cite{Fabjan82},
where $\lambda_I$ is $\lambda_{\pi}^{eff}$ in our case.
For the entire Tile calorimeter the $99\%$ containment radius is equal to
$1.7 \pm 0.1$ $\lambda_{\pi}^{eff}$.
Based on our study, we believe that it is a poor approximation to regard
the values obtained from the marginal density function
or the energy depositions in strips as the measure of the transverse
shower containment, as was done in
\cite{womersley88}.
In that paper the value of $1.1\ \lambda_{\pi}^{eff}$
was obtained for $99 \%$ containment at 100 GeV,
and the conclusion was drawn that
their ``result is consistent with the {\it rule of thumb}
that a shower is contained within a cylinder of radius equal
to the interaction length of a calorimeter material''.
However Tile calorimeter measurements show that the cylinder radius
for $99\%$ shower containment is about two interaction lengths.
If we extract the lateral shower containment dimension
using instead the integrated function
$F(z)$, given in Fig.\ \ref{fig:f9-1},
we obtain the value of $300$ mm or $1.2\ \lambda_{\pi}^{eff}$,
which agrees with \cite{womersley88}.
\section{Longitudinal Profile}
We have examined the differential deposition of energy $\Delta E/ \Delta x$
as a function of $x$, the distance along the shower axis.
Table \ref{long-1} lists
the centers in $x$ of the depth segments, $x$,
and the lengths along $x$ of the depth segments,
$\Delta x$, in units of $\lambda_{\pi}^{Fe}$,
the average shower energy depositions in various depth segments, $E_{0}$,
and the energy depositions
per interaction length
$\lambda_{\pi}^{Fe}$, $\Delta E / \Delta x$.
Note that the values of $E_0$ have been obtained taking into account
the longitudinal energy leakage which amounts to 1.8 GeV for 100 GeV
\cite{budagov96-76}.
Our values of $\Delta E / \Delta x$
together with the data of \cite{hughes90}
and Monte Carlo predictions (GEANT-FLUKA + MICAP) \cite{juste95} are
shown
in Fig.\ \ref{fig:f15}.
The longitudinal energy deposition
for our calorimeter using longitudinal orientation of the scintillating
tiles is in good agreement with that of a
conventional iron-scintillator
calorimeter.
The longitudinal profile, $\Delta E / \Delta x$, may be approximated
using two parametrizations.
The first form is
\begin{equation}
\frac{dE(x)}{dx} =
\frac{E_{f}\ \beta^{\alpha + 1}}{\Gamma (\alpha + 1)}\
x^{\alpha}\ e^{-\beta x}
\label{elong00}
\end{equation}
where $E_{f} = E_{beam}$, and $\alpha$ and $\beta$ are free parameters.
Our data at 100 GeV and those of Ref.\ \cite{hughes90} at 100 GeV were
jointly fit to this expression; the fit is shown in Fig.\ \ref{fig:f15}.
The second form is the analytical representation of the
longitudinal shower profile from the front of the calorimeter
\begin{eqnarray}
\frac{dE (x)}{dx} & = &
N\
\Biggl\{
\frac{w\ X_{0}}{a}\
\biggl( \frac{x}{X_{0}} \biggr)^a\
e^{- b \frac{x}{X_{0}}}\
{}_1F_1 \biggl( 1,\ a+1,\
\biggl(b - \frac{X_{0}}{\lambda_I} \biggr)\ \frac{x}{X_{0}}
\biggr)
\nonumber \\
& & + \
\frac{(1 - w)\ \lambda_I}{a}\
\biggl( \frac{x}{\lambda_I} \biggr)^a\
e^{- d \frac{x}{\lambda_I}}\
{}_1F_1 \biggl( 1,\ a+1,\
\bigl( d - 1 \bigr)\ \frac{x}{\lambda_I} \biggr)
\Biggr\} ,
\label{elong2}
\end{eqnarray}
where ${}_1F_1$ is the confluent hypergeometric function
\cite{abramovitz64}. Here the depth variable, $x$, is the depth in
equivalent $Fe$, $X_0$ is the radiation length in $Fe$
and in this case $\lambda_I$ is $\lambda_\pi^{Fe}$.
The normalisation factor, $N$, is given by
\begin{equation}
N =
E_{beam} \biggl/ \int \limits_{0}^{\infty} \frac{dE(x)}{dx} dx
= \frac{E_{beam}}{\lambda_I\ \Gamma (a)\
\bigl(w\ X_{0}\ b^{-a} + (1-w)\ \lambda_I\ d^{-a} \bigr)
} .
\label{elong02}
\end{equation}
This form was suggested in \cite{kulchitsky98} and derived by integration
over the shower vertex positions of the longitudinal shower development
from the shower origin
\begin{equation}
\frac{dE (x)}{dx} =
\int \limits_{0}^{x}
\frac{dE_{s}(x-x_v)}{dx}\ e^{- \frac{x_v}{\lambda_I}}\ dx_v ,
\label{elong1}
\end{equation}
where $x_v$ is the coordinate of the shower vertex.
(This is necessary because with the Tile calorimeter
longitudinal segmentation the shower vertex is not measured).
For the para\-me\-trization of longitudinal shower development,
the well known parametrization
suggested by Bock et al.\ \cite{bock81} has been used
\begin{equation}
\frac{dE_{s} (x)}{dx} =
N\
\Biggl\{
w\ \biggl( \frac{x}{X_{0}} \biggr)^{a-1}\
e^{- b \frac{x}{X_{0}}}\
+ \
(1-w)\
\biggl( \frac{x}{\lambda_I} \biggr)^{a-1}\
e^{- d \frac{x}{\lambda_I}}
\Biggr\},
\label{elong0}
\end{equation}
where $a,\ b,\ d,\ w$ are parameters.
We compare the form (\ref{elong2})
to the experimental points at 100 GeV using the
parameters calculated in Refs.\ \cite{bock81} and \cite{hughes90}. Note
that now we are not performing a fit but checking how well the general
form (\ref{elong2})
together with two sets of parameters for iron-scintillator
calorimeters describe our data.
As shown in Fig.\ \ref{fig:f15}, both sets of
parameters work rather well in describing the 100 GeV data.
Turning next to the longitudinal shower development at different
energies, in Fig.\ \ref{fig:f15-l} our values of $\Delta E / \Delta x$
for 20 -- 300 GeV together with the data from \cite{hughes90} are shown.
The solid and dashed lines are calculations with function (\ref{elong2})
using parameters from \cite{hughes90} and \cite{bock81}, respectively.
Again, we observe reasonable agreement between our data and the
corresponding data for conventional iron-scintillator
calorimeter on one hand, and between data and the
parametrizations described above.
Note that the fit in \cite{hughes90} has been performed in the energy
range from 10 to 140 GeV; hence
the curves for 200 and 300 GeV should be considered as extrapolations.
It is not too surprising that at these energies the agreement is
significantly worse, particularly at 300 GeV.
In contrast, the parameters of \cite{bock81}
were derived from data spanning the range 15 -- 400 GeV, and are in much
closer agreement with our data.
\section{The parametrization of Hadronic Showers}
The three-dimensional parametrization for spatial hadronic shower
development is
\begin{equation}
\Psi (x, r) =
\frac{dE(x)}{dx} \cdot
\frac{
\sum \limits_{i=1}^{3} \ \frac{a_{i} (x)}{\lambda_{i} (x)}\
K_{0} \bigl( \frac{r}{\lambda_{i} (x)} \bigr)
}{
2\pi\ \sum \limits_{i=1}^{3} a_{i}(x) \lambda_{i} (x)} ,
\label{e123}
\end{equation}
where $dE (x)/ dx$, defined by equation (\ref{elong2}),
is the longitudinal energy deposition,
the functions $a_{i}(x)$ and $\lambda_{i} (x)$ are given
by equations (\ref{eai}) and (\ref{eli}), and
$K_{0}$ is the modified Bessel function.
This explicit three dimensional parametrization can be used
as a convenient tool for many calorimetry problems requiring
the integration of a shower energy deposition in a volume
and the reconstruction of the shower coordinates.
\section{Electromagnetic Fraction of Hadronic Showers}
One of the important issues in the understanding of hadronic showers
is the electromagnetic component of the shower, i.\ e.\ the fraction
of energy going into $\pi^0$ production and its dependence on radial
and longitudinal coordinates, $f_{\pi^{0}} (r, x)$.
Following \cite{acosta92},
we assume that the electromagnetic part of a hadronic shower is the
prominent central core, which in our case is the first term in the
expression (\ref{e23}) for the radial energy density function, $\Phi(r)$.
Integrating $f_{\pi^0}$ over $r$ we get
\begin{equation}
\label{e231}
f_{\pi^{0}} =
\frac{a_1 \lambda_1}{\sum \limits_{i=1}^{3} a_i \lambda_i} .
\end{equation}
For the entire Tile calorimeter this value is $(53 \pm 3) \%$
at 100 GeV.
The observed $\pi^{0}$ fraction, $f_{\pi^{0}}$, is related to the
intrinsic actual fraction, $f^{\prime}_{\pi^{0}}$, by the equation
\begin{equation}
f_{\pi^{0}}(E) =
\frac{e\ E^{\prime}_{em}}{e\ E^{\prime}_{em} + h\ E^{\prime}_{h}} =
\frac{ e/h \cdot f^{\prime}_{\pi^{0}}(E)}{(e/h - 1)
\cdot f^{\prime}_{\pi^{0}}(E) + 1} ,
\label{e506-1}
\end{equation}
where $E^{\prime}_{em}$ and $E^{\prime}_{h}$ are the intrinsic
electromagnetic and hadronic parts of shower energy,
$e$ and $h$ are the coefficients of conversion of intrinsic electromagnetic
and hadronic energies into observable signals,
$f^{\prime}_{\pi^{0}} = E^{\prime}_{em}/(E^{\prime}_{em} + E^{\prime}_{h})$.
There are two analytic forms for the intrinsic $\pi^{0}$
fraction suggested by Groom
\cite{groom90}
\begin{equation}
f^{\prime}_{\pi^{0}}(E) =
1 - { \biggl( \frac{E}{E_{0}^{\prime}} \biggr) }^{m-1}
\label{e506-2}
\end{equation}
and Wigmans \cite{wigmans88}
\begin{equation}
f^{\prime}_{\pi^{0}}(E) =
k \cdot ln \biggl( {\frac{E}{E_{0}^{\prime}}} \biggr),
\label{e506-3}
\end{equation}
where $E_{0}^{\prime} = 1$ GeV, $m = 0.85$ and $k = 0.11$.
We calculated $f_{\pi^{0}}$ using the value $e/h = 1.34\pm0.03$
for our calorimeter
\cite{tilecal-tdr96}, \cite{budagov95-72}
and obtained the curves shown in Fig.\ \ref{fig:f150-1}.
Our result at 100 GeV is compared in
Fig.\ \ref{fig:f150-1}
to the modified Groom and Wigmans parametrizations and to results from
the Monte Carlo codes
CALOR \cite{gabriel94},
GEANT-GEISHA \cite{juste95} and GEANT-CALOR \cite{bosman97}
(the latter code is an implementation of
CALOR89 differing from GEANT-FLUKA only for
hadronic interactions below 10 GeV).
Note that the Monte Carlo calculations were performed for the intrinsic
$\pi^{0}$ fraction, $f^{\prime}_{\pi^{0}}(E)$, and therefore
the results were modified by us according to (\ref{e506-1}).
As can be seen from Fig.\ \ref{fig:f150-1}, our calculated
value of $f_{\pi^{0}}$ is about one standard deviation lower than
two of the Monte Carlo results
and the Groom and Wigmans parametrizations.
Figure \ref{fig:f150-3} shows the fractions $f_{\pi^{0}} (r)$ as a
function of $r$.
As can be seen, the fractions $f_{\pi^{0}} (r)$ for the entire calorimeter
and for depth segments 1 -- 3 amount to about $90\%$ as $r \rightarrow 0$
and
decrease to about $1\%$ as $r \rightarrow \lambda_{\pi}^{eff}$.
However for depth segment 4 the value of $f_{\pi^{0}} (r)$ amounts
to only $50\%$ as $r \rightarrow 0$ and decreases slowly to about $10\%$ as
$r \rightarrow \lambda_{\pi}^{eff}$.
Figure\ \ref{fig:f150-2}
shows the values of $f_{\pi^{0}}(x)$ as a function of $x$,
as well as the linear fit which gives
$f_{\pi^{0}}(x) = (75\pm2) - (8.4\pm0.4) x$ ($\%$).
Using the values of $f_{\pi^{0}}(x)$ and energy depositions for various
depth segments, we obtained the contributions from the electromagnetic and
hadronic parts of hadronic showers in Fig.\ \ref{fig:f15}.
The curves represent a fit to the electromagnetic
and hadronic components of the shower using equation (\ref{elong00}).
$E_{f}$ is set equal to $f_{\pi^{0}} E_{beam}$
for the electromagnetic fraction and
$(1 - f_{\pi^{0}}) E_{beam}$ for the hadronic fraction.
The electromagnetic component of a hadronic shower rise and decrease
more rapidly than the hadronic one
($\alpha_{em} = 1.4\pm0.1$, $\alpha_{h} = 1.1\pm0.1$,
$\beta_{em} = 1.12\pm0.04$, $\beta_{h} = 0.65\pm0.05$).
The shower maximum position
($x_{max} = (\alpha / \beta )\ \lambda_{\pi}^{eff}$)
occurs at a shorter distance from the calorimeter front face
($x_{max}^{em} = 1.23\ \lambda_{\pi}^{eff}$,
$x_{max}^{h} = 1.85\ \lambda_{\pi}^{eff}$).
At depth segments greater than $4\ \lambda_{\pi}^{eff}$, the hadronic
fraction of the shower begins to dominate.
This is natural since
the energy of the secondary hadrons is too low to permit significant
pion production.
\section{Summary and Conclusions}
We have investigated the lateral development of hadronic showers
using 100 GeV pion beam data at an incidence angle of
$\Theta = 10^{\circ}$ for impact points $z$ in the range from
$- 360$ to $200$ mm
and the longitudinal development of hadronic showers
using 20 -- 300 GeV pion beams at an incidence angle of
$\Theta = 20^{\circ}$.
Some useful formulae for the investigation of lateral profiles
have been derived using a three-exponential form of
the marginal density function $f(z)$.
We have obtained for four depth segments and for the entire calorimeter:
energy depositions in towers, $E(z)$;
cumulative functions, $F(z)$;
underlying radial energy densities, $\Phi(r)$;
the contained fraction of a shower as a function of radius, $I(r)$;
the radii of cylinders for a given shower containment fraction;
the fractions of the electromagnetic and hadronic parts of a shower;
differential longitudinal energy deposition $\Delta E / \Delta x$;
and a three-dimensional hadronic shower parametrization.
We have compared our data with those from a conventional
iron-scintilla\-tor ca\-lo\-ri\-me\-ter, those from a lead-scintillator
fiber calorimeter, and with Monte Carlo calculations.
We have found that there is general reasonable agreement in the
behaviour of the Tile calorimeter radial density functions
and those of the lead-scintillating fiber calorimeter;
that the longitudinal profile agrees with that of a conventional
iron-scintillator calorimeter and the Monte Carlo predictions;
that the value at 100 GeV
of the calculated
fraction of energy going into $\pi^{0}$
production in a hadronic shower, $f_{\pi^{0}}$, agrees with
the Groom and Wigmans parametrizations and with
some of the Monte Carlo predictions.
The three-dimensional parametrization of hadronic showers
that we obtained allows direct use in any application that requires
volume integration of shower energy depositions
and position reconstruction.
The experimental data on the transverse and longitudinal
profiles, the radial energy densities and the
three-dimensional hadronic shower parametrization are
useful for understanding the performance of the Tile calorimeter,
but might find broader application
in Monte Carlo modeling of hadronic showers, in particular in
fast simulations, and for future calorimeter design.
\section{Acknowledgements}
This paper is the result of the efforts of many people from the ATLAS
Collaboration.
The authors are greatly indebted to the entire Collaboration
for their test beam setup and data taking.
We are grateful to the staff of the SPS, and in
particular to Konrad Elsener, for the excellent beam conditions and
assistance provided during our tests.
|
\section{Introduction}
The observed large-scale structure in the Universe has been currently
addressed, within the framework of gravitational instability, by two
families of models: initial density perturbations can either be due to
``freezing in'' of quantum fluctuations of a scalar field (inflaton)
during an inflationary era~\cite{infl}, or they may be seeded by a
class of topological defects, naturally formed during a
symmetry-breaking phase transition in the early Universe~\cite{kibble}. The
recent bulk of observational and experimental data and, in particular,
the cosmic microwave background anisotropy measurements, and the
redshift surveys of the distribution of galaxies and clusters of
galaxies, impose severe constraints on the two families of models, as
well as on the variety of possible scenarios introduced within each
family.
\par
The simplest topological defects models of structure
formation show conflicts with observational data. As was first
shown in Ref.~\cite{ram}, global topological defects models predict
strongly suppressed acoustic peaks. While on large angular scales the
predicted cosmic microwave background radiation (CMBR) spectrum is in
good agreement with COBE measurements, on smaller angular scales the
topological defects models cannot reproduce the data of the Saskatoon
experiment. One can manufacture models~\cite{ruthmairi} with
structure formation being induced by scaling seeds, which lead to an
angular power spectrum with the same characteristics (position and
amplitude of acoustic peaks), as the one predicted by standard
inflationary models. The open question is, though, whether such models
are the outcome of a realistic theory. However, the most severe
problem for topological defects models of structure formation is their
predicted~\cite{pst}, ~\cite{abr} lack of large-scale power in the matter power
spectrum, once normalized to COBE.
Choosing scales of $100 h^{-1}{\rm Mpc}$, which are most probably
unaffected by non-linear gravitational evolution, standard topological
defect models, once normalized to COBE, require a bias factor
($b_{100}$) on scales of $100 h^{-1}{\rm Mpc}$ of $b_{100}\approx 5$, to
reconcile the predictions for the density field fluctuations with the
observed galaxy distribution. However, the latest theoretical and
experimental studies favour a current value of $b_{100}$ close
to unity.
\par
In what follows, we shall place ourselves within the framework of
cosmological perturbations of quantum-mechanical origin in the context
of inflationary models. The inflationary paradigm was proposed in
order to explain the shortcomings of the standard (Big Bang)
cosmological model. In addition, it offers a scenario for the
generation of the primordial density perturbations, which can lead to
the formation of the observed large-scale structure.
\par
The theory of cosmological perturbations of quantum-mechanical origin
rests on two well-established theories. On the one hand, (linearized)
general relativity allows a calculation of the evolution and the
amplification of perturbations throughout the cosmic evolution; the
mechanism at work being parametric amplification of the fluctuations
due to the interaction of the perturbations with the
background~\cite{Grireview}. On the other hand, quantum field theory
permits to understand the origin of these perturbations. If the
quantum fields are initially, i.e. at the beginning of inflation,
placed in the vacuum state, then because of the Heisenberg principle,
fluctuations are unavoidable. Moreover, the amplitude of these
fluctuations is completely fixed.
\par
Inflation, employing the theory of cosmological perturbations of
quantum-mechanical origin, leads to definite predictions for the
anisotropies of the CMBR, as well as for the power spectrum, which can
be tested against experimental and observational data. In particular,
simple models predicts a scale-invariant spectrum, with, provided the quantum
fields are initially placed in the vacuum, Gaussian fluctuations.
\par
Let us briefly discuss the observational data, namely the CMBR
anisotropies measurements and the redshift
surveys of the distribution of galaxies.
\par
The CMBR, last scattered at the epoch
of decoupling, has to a high accuracy a black-body distribution
\cite{bb}, with a temperature $T_0=2.728\pm 0.002 ~{\rm K}$, which is
almost independent of direction. The DMR experiment on the COBE
satellite measured a tiny variation in intensity of the CMBR, at fixed
frequency. This is equivalently expressed as a variation ${\rm \delta
}T$ in the temperature, which was measured to be
${\rm \delta }T/T_0 \approx 10^{-5}$~\cite{cobe}.
The 4-year COBE data are fitted by a scale-free
spectrum; the spectral index was found to be $n_{\rm S}=1.2\pm 0.3$
and the quadrupole anisotropy $Q_{\rm rms-PS}=15.3^{+3.8}_{-2.8} ~{\rm
\mu K}$~\cite{cobe}. The CMBR anisotropies spectrum is usually
parametrized in terms of the multipole moments $C_\ell$, defined as
the coefficients in the expansion of the temperature autocorrelation
function
\begin{eqnarray}
\label{C_l}
\left\langle{{\rm \delta} T\over T}({\bf e}_1)
{{\rm \delta} T\over T}({\bf e}_2)\right\rangle
\bigg|_{({\bf e}_1\cdot {\bf e}_2=\cos\vartheta)} =
\nonumber \\
{1\over 4\pi}\sum_\ell (2\ell & + & 1)\,C_\ell P_\ell(\cos\vartheta)\;~,
\end{eqnarray}
which compares points in the sky separated by an angle $\vartheta$.
The value of $C_\ell$ is determined by fluctuations on angular scales
of order $\pi/\ell$. The angular power spectrum of anisotropies
observed today is usually given by the power per logarithmic interval
in $\ell$, plotting $\ell(\ell+1)C_\ell$ versus $\ell$.
\par
On large angular scales, the main physical mechanism which contributes
to the redshift of photons propagating in a perturbed Friedmann
geometry, originates from fluctuations in the gravitational potential
on the last-scattering surface. The COBE-DMR experiment, which
measured CMBR anisotropies on such large angular scales ($\ell ~ \lower2pt\hbox{$\buildrel{<}\over{\sim}$} 20$),
confirmed the predicted scale-invariant spectrum and yields
mainly a normalization for the different
models of large-scale structure formation.
\par
On intermediate angular scales, $0.1^\circ\stackrel{<}{\sim}
\vartheta\stackrel{<}{\sim} 2^\circ$, the main contribution to the
CMBR anisotropies comes from the intrinsic inhomogeneities on the
surface of the last scattering, due to acoustic oscillations in the
coupled baryon-radiation fluid prior to decoupling. On the same
angular scales, there is a Doppler contribution to the CMBR
anisotropies, due to the relative motions of emitter and observer. The
sum of these two contributions is denoted by the term acoustic peaks.
An analysis of recent CMBR flat-band measurements on intermediate
angular scales gives~\cite{charley} in the best-fit power spectrum a
peak $[\ell(\ell+1)C_\ell/2\pi]^{1/2}T_0= 76 ~{\rm \mu K}$ with
$\ell=260$.
\par
Among the various experiments measuring CMBR anisotropies, the
Saskatoon experiment~\cite{sask} is of particular importance since it
relates~\cite{einasto} CMBR anisotropies to the power spectrum of
matter density perturbations estimated through clustering properties
of galaxies and clusters of galaxies. More precisely, the Saskatoon
experiment measures temperature anisotropies for multipoles in the
range $\ell \approx 80 - 400$, which corresponds to the range of
wavelengths for which we have data on galaxy clusters.
\par
Analysing a large number of available data on redshifts of individual
galaxies and Abell galaxy clusters, one obtains~\cite{einasto} the
power spectrum for clusters of galaxies, over the wave-number interval
from $k\approx 0.03 ~h$~Mpc${}^{-1}$ to $k\approx 0.3 ~h~$Mpc${}^{-1}$. On
very large scales ($k<0.03 ~h$~Mpc${}^{-1}$), the large error bars
are due to incomplete data. However, near the turn over, error bars are
small, thus both the relative position and amplitude of the turn over
are determined accurately. As discussed in, e.g. Ref.~\cite{einasto},
the power spectrum reveals the existence of a non-trivial feature at a
wave-number $k_0 =0.052\pm0.005 ~h~{\rm Mpc}^{-1}$. Assuming this peak exists
(further studies are necessary to confirm it), the amplitude
of the observed power spectrum is larger near the peak by a factor
$1.4$ \cite{einasto} with respect to the power spectrum of the
standard cold dark matter model. The existence of this peak is not
related~\cite{einasto2} to acoustic oscillations in the tight coupled
baryon--photon plasma. As stated in Ref.~\cite{einasto2}, the current
CMBR experimental data combined with observational cluster data,
favour theoretical models that have built-in a characteristic scale
in their initial spectrum.
\par
Recently, the COBE data have also been used to test the gaussianity of
the CMBR anisotropies. Three
groups~\cite{non-gaus1,non-gaus2,non-gaus3} have now reported results
showing that the fluctuations would not be Gaussian. The three groups
work with different methods. In Ref.~\cite{non-gaus1}, the estimation
of the bispectrum $B_\ell$ is used as a criterion to test
gaussianity. The dominant non-Gaussian contribution has been found
near $\ell=16$. It is clear that these results should be taken
cautiously since, for example, the issue of foreground contamination
could change the conclusions. However, the possibility of non-Gaussian
statistics in the CMBR anisotropies should be taken into
account seriously.
\par
The Saskatoon measurements could be explained by playing with the
values of the cosmological parameters. In particular, the value of the
cosmological constant, $\Omega _{\Lambda }\approx 0.6$, recently
inferred from the SNIa measurements could account for the high position
of the first acoustic peak. But the other features (the presence of a
peak in the power spectrum; non-Gaussian fluctuations in the CMBR), if
confirmed, clearly go beyond the paradigm of cold dark matter (CDM)
and slow roll inflation. In order to explain them, different
mechanisms have been advocated. For instance, double-inflation
\cite{di} or multiple-inflation~\cite{mi} models have been used to
explain the presence of the peak in the power spectrum. Another
scenario can be offered within models where the inflaton field evolves
through a kink in its potential~\cite{kink}. To explain the
non-Gaussianity, different mechanisms have been proposed. Of course, this
appears naturally when the perturbations are induced by topological
defects. However, even in the context of inflation, non-Gaussianity
can be present, as for example in the case of stochastic
inflation~\cite{Alejandro,stochastic}.
\par
All these solutions are in fact different possible modifications of
the power spectrum of the primordial fluctuations. In this article,
our aim is to discuss the choice of the initial quantum state in which
the quantum fields are placed. This choice is of course crucial for
the determination of the primordial power spectrum, and different
quantum states will lead to different power spectra. In the
literature, it is (almost, see Ref.~\cite{Staro}) always assumed that
the state of the perturbations is the vacuum (In a curved spacetime
the definition of the vacuum state is not unique. A more precise
definition of the vacuum used in this paper is given in what follows
and coincides with the one in Ref.~\cite{rbv}):
\begin{equation}
\label{defvacuum}
|0\rangle \equiv \bigotimes _{\bf k} |0_{\bf k}\rangle .
\end{equation}
Let us examine how this choice can be justified. Since this question
is a problem of boundary conditions, it must be addressed by means of
a theory of the initial conditions for the early Universe. Such a
theory should rely on full quantum gravity, which is
unknown at present. The only candidate at our disposal is quantum
cosmology. Generally, it predicts that the initial state is indeed the
vacuum. For example, the no-boundary choice for the wave function of
the cosmological perturbations implies that the Bardeen operator is
placed in the vacuum state, see Ref.~\cite{HHalli}. This result does
not come as a surprise since the Hartle--Hawking proposal is a
generalization of a method that gives the ground-state wave function
of a system in ordinary quantum mechanics.
\par
However, although fascinating, quantum cosmology is not yet a
well-developed branch of physics and many important questions remain
unsolved to this day. To our knowledge, there exists no proof that
quantum considerations automatically lead to a vacuum initial state
for the perturbations. Such a proof, if it exists, should rely on full
quantum gravity.
\par
On the other hand, the choice of the vacuum is also based on the
hypothesis that the initial state of the Universe should be a
``maximally symmetric state''~\cite{mssstaro}. Concretely, this means
that no scale should be privileged. This seems to be the simplest
starting point. However, since the choice of the initial state is
supposed to appear naturally in the context of quantum gravity, it
could also be argued that such a privileged scale does exist and is equal
to the Planck scale, $l_{\rm Pl}=(\hbar G/c^3)^{1/2}
\approx 10^{-33} ~$cm. This becomes even more intriguing if one recalls
that in order to solve the usual problems of the standard model of
cosmology, one needs $60$ $e$-folds during inflation. This means that
the Planck scale has now been stretched to a scale of at least $60$
pc. Accordingly, all the wavelengths below $60$ pc were sub-Planckian
at the moment of their generation. Of course, the structure of
space-time below the Planck scale is unknown and it may be very
different from the one we are used to. Probably, such notions as
sub-Planckian wavelengths or even scale factor are meaningless in a
regime where the gravitational quantum effects are important.
\par
The arguments presented in the previous discussion show that it is
worth studying non-vacuum initial states for cosmological
perturbations. Rather than relying on theoretical arguments, our goal
will be to allow for the possibility of non-vacuum initial states and
to establish the consequences for the observables described at the
beginning of the introduction. We study whether choices other than the
vacuum automatically lead to inconsistencies or conflicts with the
available observations or if, on the contrary, there exists a window for
the free parameters of the model, which fits the observational data.
\par
Our choice of a non-vacuum initial state is guided by a very simple
idea: the initial state could have a built-in characteristic
scale since this seems to be the simplest way to generalize the vacuum
state. The question now arises as for the physical origin of
this scale. A possible answer is that the natural scale is the Planck
length stretched by the cosmological expansion. It is clear that, so as not to
be in conflict with
observations, we would like this fundamental scale to be now
translated to the characteristic scale $l_{\rm C}\approx
200\,{\rm Mpc}=6.2\cdot 10^{26}$cm (here, as in the rest of this article we
take $h=0.5$). Since the ratio $l_{\rm C}/l_{\rm Pl}$ is given by
$l_{\rm C}/l_{\rm Pl}\approx 10^{27}e^N$, where $N$ is the number of
$e$-folds during inflation, this means that $N\approx
75$. Interestingly enough, we note that this leads to a number of
$e$-folds greater than the minimum number required, i.e. $60$. In the
context of Linde's chaotic inflation, it is assumed that, initially,
the inflaton potential $V(\varphi )$ is such
that $V(\varphi _i) \approx m_{\rm Pl}^4$ where $m_{\rm Pl}$ is the
Planck mass. If the potential is given by
e.g. $V(\varphi )=(\lambda /4!)\varphi ^4$, this leads to an initial value of the
scalar field greater than $4.4m_{\rm Pl}$, which is needed to get the usual $60$
$e$-folds. Consequently, this leads to a huge number of $e$-folds, $N\approx 10^8$. It
is clear that, with such a number, the Planck length cannot be stretched to
$200\, {\rm Mpc}$ presently. Let us note that these models (with a large
number of $e$-folds) suffer from the ``super-Planck scale
problem''~\cite{Robert}: all the scales of cosmological interest now
were sub-Planckian at the beginning of inflation. Since quantum field theory
is expected to break down in this regime, the predictions of these
models could be questionable. On the other hand, in the spirit of chaotic
inflation itself, there exists regions of space in which the initial value of the
field was $\varphi _i\approx 4.9m_{\rm Pl}$. This value leads to a number
of $e$-folds equal to $75$. Therefore, the model presented in this article
is certainly more relevant in the case where inflation does not last for a
long period. In chaotic inflation, the probability of having a long period
of inflation is greater than the probability to get a small
number of $e$-folds. Thus, our model does not fit very well within the
chaotic
inflation approach.
\par
It should also be mentioned that it has been shown in
Ref.~\cite{Robertinistate} that a large class of initial states approaches
the Bunch Davis vacuum in the de Sitter spacetime. However, this class of initial
states has a Gaussian wave functional and therefore the argument that the choice
of a non-vacuum initial state would involve exponential fine tunning does not apply
to the case considered here.
\par
Recently, a model with a small number of $e$-folds has been constructed
in Ref.~\cite{Sarkar}. This kind of models naturally arises in the context
of supersymetric (SUSY and SUGRA) inflation. They are particulary well
suited to the model put forward in this article. They consist in
multiple bursts of inflation which in total last for $\approx 75$
expansion times. In addition, the last stage of inflation is preceded
by other inflationary epochs. The ``initial state'' of this last
epoch is the result of the evolution of the ``true initial state'' through
the multiple preceding bursts of inflation. Clearly, there is no
reason for assuming this ``effective initial state'' to be the vacuum. Let
us emphasize that this argument holds for every model with many stages
of inflation since, in this case, the origin of the characteristic
scale could no longer be the Planck length stretched to $l_{\rm C}$ but
could correspond to the time where one of the fields starts rolling down.
\par
Our model possesses a privileged scale and therefore belongs to the
class of models already envisaged in Ref.~\cite{bsi}. However, we
would like to emphasize that the origin of this scale is physically
completely different and we will point out that there exist
observables, which, in principle, allow us to distinguish between the
different models. A last comment on the fine tuning issue is in
order here. It is true that the position of the characteristic scale must
be chosen carefully. Otherwise the model would simply
be in contradiction with the available data. We would like to
emphasize that this is not a feature of our model only but in fact
of all the BSI (broken scale invariant) models~\cite{bsi}. In this
respect our model is similar to the other BSI models.
\par
This paper is organized as follows: In Section II we discuss non-vacuum
initial state for the cosmological perturbations. We first
briefly describe the theory of perturbations of quantum-mechanical
origin. We then describe the non-vacuum initial states considered in
this article. We finally calculate the power spectra of the Bardeen
potential for these states and show that it possesses either
a step or a bump. In Section III we examine the
observational consequences of the calculated power spectra; we compare
our theoretical predictions with current experimental and
observational data, which will fix the parameters of our model. We end
with the conclusions given in Section IV.
\section{Non-vacuum initial state for the perturbations}
\subsection{Perturbations of quantum-mechanical origin}
The background model is described by a
Friedmann--Lema\^{\i}tre--Robertson--Walker (FLRW) metric whose space-like sections are flat
($c=1$): ${\rm d}s^2=a^2(\eta )(-{\rm d}\eta ^2 +{\rm d}{\bf
x}^2)$. We assume that inflation is driven by a single scalar field,
$\varphi _0(\eta )$. It is convenient to define the background
quantities ${\cal H}(\eta )$ and $\gamma (\eta )$ by:
\begin{equation}
\label{defH,gamma}
{\cal H}\equiv a'/a, \quad \gamma (\eta )
\equiv 1-\frac{{\cal H}'}{{\cal H}^2},
\end{equation}
where the primes denote the derivatives with respect to conformal time. In the
case of the de Sitter space-time, $\gamma $ vanishes.
\par
In the synchronous gauge, without loss of generality, the line element
for the FLRW background plus scalar perturbations can be written
as~\cite{Griden}:
\begin{eqnarray}
\label{metricsg}
{\rm d}s^2 &=& a^2(\eta )\biggl\{-{\rm d}\eta ^2+\biggl[\delta _{ij}
+\frac{1}{(2\pi )^{3/2}}
\int {\rm d}{\bf k}\biggl(h(\eta ,{\bf k})\delta _{ij}\nonumber \\
& &-\frac{h_l(\eta ,{\bf k})}{k^2}k_ik_j\biggr)
e^{i{\bf k}\cdot{\bf x}}\biggr]{\rm d}x^i{\rm d}x^j\biggr\},
\end{eqnarray}
where the functions $h, h_l$ represent the scalar perturbations of the
gravitational field and the longitudinal--longitudinal perturbation,
respectively. In the same manner, the perturbations of the scalar
field are Fourier decomposed according to:
\begin{equation}
\label{defdeltasf}
{\rm \delta }\varphi (\eta ,{\bf x})=\frac{1}{(2\pi )^{3/2}}
\int {\rm d}{\bf k}\varphi _1(\eta ,{\bf k})e^{i{\bf k}\cdot {\bf x}}.
\end{equation}
The perturbed Einstein equations couple the scalar sector, $h$ and
$h_l$, to the perturbed scalar field $\varphi _1$: see
Ref.~\cite{Griden}. It can be shown that the residual gauge invariant
quantity $\mu (\eta ,{\bf k})$
\cite{Griden} defined by:
\begin{equation}
\label{defmu}
\mu \equiv \frac{a}{{\cal H}\sqrt{\gamma }}(h'+{\cal H}\gamma h),
\end{equation}
can be used to express all other relevant quantities. The quantity $\mu
(\eta ,{\bf k})$ is related to the gauge-invariant Bardeen potential
through the equation:
\begin{equation}
\label{RPhimu}
\Phi ^{({\rm SG})}=\frac{{\cal H}\gamma }{2k^2}
\biggl(\frac{\mu }{a\sqrt{\gamma }}\biggr)',
\end{equation}
where ``{\rm SG}'' means calculated in the synchronous gauge, see
Ref.~\cite{MS1}. The quantity $\mu (\eta ,{\bf k})$ is not defined in
the de Sitter case. This case must be treated separately and for it we
have $\Phi ^{({\rm SG})}=0$: there are no density perturbations at all
because the fluctuations of the scalar field are not coupled to the
perturbations of the metric when the equation of state is
$p=-\rho$. The perturbed Einstein equations imply that the equation of
motion for $\mu (\eta ,{\bf k})$ is given by~\cite{Griden}:
\begin{equation}
\label{paraeq}
\mu ''+\biggl[k^2-\frac{(a\sqrt{\gamma })''}{(a\sqrt{\gamma })}\biggr]\mu=0.
\end{equation}
The above is the characteristic equation of a parametric
oscillator whose time-dependent frequency depends on the scale factor
and its derivative (up to $a^{(4)}$).
\par
In this article, we assume that the perturbations are of
quantum-mechanical origin. This hypothesis fixes completely the
normalization of ${\rm \delta }\varphi (\eta ,{\bf x})$ and of the
scalar perturbations. The normalization is fixed in the high-frequency
regime. In this regime, the perturbed field can be considered as a
free field in the curved FLRW background space-time. It is therefore
necessary to study the quantization of such a free field denoted in
the following by $\varphi (\eta ,{\bf x})$. We first address this
question and we then make the link between $\varphi (\eta ,{\bf x})$
and ${\rm \delta }\varphi (\eta ,{\bf x})$. We choose to define the
Fourier component of $\varphi (\eta ,{\bf x})$ according to:
\begin{equation}
\label{defchi}
\varphi (\eta ,{\bf x})=\frac{1}{a(\eta )}\frac{1}{(2\pi )^{3/2}}
\int {\rm d}{\bf k}\chi (\eta ,{\bf k})e^{i{\bf k}\cdot {\bf x}},
\end{equation}
where we have renormalized the time-dependent amplitude with the scale
factor. The Fourier component satisfies $\chi (\eta ,-{\bf k})=\chi
^*(\eta ,{\bf k})$, because the field is real. The action, given by:
\begin{eqnarray}
\label{action1}
S &=&\int {\rm d}^4x {\cal L} \\
\label{action2}
&=&\frac{1}{2}\int {\rm d}^4x
a^2\biggl[\left(\varphi ' (\eta ,{\bf x})\right)^2
-\sum _i \left({\rm \partial}_i \varphi (\eta ,{\bf x})\right)^2\biggr],
\end{eqnarray}
can also be written in terms of the Fourier component $\chi (\eta
,{\bf k})$. The result reads:
\begin{eqnarray}
\label{actionfourier}
S &=& \int {\rm d}\eta \int _{R^{3+}}{\rm d}{\bf k} \bar{{\cal L}} \\
\label{actionfourier2}
&=& \int {\rm d}\eta \int _{R^{3+}}{\rm d}{\bf k}
\biggl\{|\chi '(\eta ,{\bf k})|^2 +
\biggl(\frac{a'^2}{a^2}-k^2\biggr)|\chi (\eta ,{\bf k})|^2
\nonumber \\
& &-\frac{a'}{a}
\biggl[\chi '(\eta ,{\bf k})\chi ^*(\eta ,{\bf k})
+\chi '^{*}(\eta ,{\bf k})\chi (\eta ,{\bf k})\biggr]\biggr\}.
\end{eqnarray}
The variation of the action leads to the equation of motion for the
Fourier component: $\chi ''+\chi [k^2-a''/a]=0$. Again, we find a
parametric oscillator-type equation. Of course, if there is no
expansion, or if the Universe is in the radiation-dominated era,
it reduces to the usual equation of motion of an harmonic oscillator.
\par
We now turn to the Hamiltonian formalism. The first step is the
calculation of the momentum conjugate to $\varphi (\eta ,{\bf x})$
defined by:
\begin{equation}
\label{defpi}
\pi (\eta, {\bf x})\equiv \frac{{\rm \partial}{\cal L}}
{{\rm \partial}(\varphi '(\eta ,{\bf x}))}=a^2\varphi '(\eta ,{\bf x}).
\end{equation}
$\pi (\eta ,{\bf x})$ can be expressed in terms of the momentum
conjugate to $\chi (\eta ,{\bf k})$,
\begin{equation}
\label{defp}
p(\eta, {\bf k})\equiv\frac{{\rm \partial}{\bar{\cal L}}}{{\rm \partial}(
\chi '^*(\eta ,{\bf k}))}
=\chi '(\eta ,{\bf k})-\frac{a'}{a}\chi (\eta ,{\bf k}),
\end{equation}
through the relation:
\begin{equation}
\label{Rpip}
\pi (\eta, {\bf x})=\frac{a(\eta )}{(2\pi)^{3/2}}\int {\rm d}{\bf k}
p(\eta, {\bf k})e^{i{\bf k}\cdot {\bf x}}.
\end{equation}
As a preparation to the quantization, the normal variable $\alpha
(\eta ,{\bf k})$ is introduced. Its definition is given by:
\begin{equation}
\label{alpha}
\alpha (\eta ,{\bf k})\equiv
N(k)\biggl[\chi (\eta ,{\bf k})+\frac{i}{k}p(\eta ,{\bf k})\biggr],
\end{equation}
where $N(k)$ is, at the classical level, a free factor.
\par
We are now in a position where quantization can be carried out. The
field $\varphi (\eta ,{\bf x})$ and its conjugate momentum $\pi (\eta
,{\bf x})$ become operators that satisfy the commutation relation:
\begin{equation}
\label{commutator}
[\hat{\varphi }(\eta ,{\bf x}),\hat{\pi }(\eta, {\bf x}')]
=i\hbar\delta ({\bf x}-{\bf x}').
\end{equation}
The normal variable $\alpha (\eta ,{\bf k})$ becomes a dimensionless
operator $c_{\bf k}(\eta )$ such that, at any time, $[c_{\bf k}(\eta
), c_{{\bf k}'}^{\dagger }(\eta )] =\delta_{{\bf k}{\bf k}'}$. With
the help of Eqs.~(\ref{defchi}) and (\ref{Rpip}) and of the definition
of the normal variable, Eq.~(\ref{alpha}), the field operator and the
conjugate momentum operator can now be expressed in terms of the
annihilation and creation operators $c_{\bf k}(\eta )$ and $c_{{\bf
k}}^{\dagger }(\eta )$. The normalization factor $N(k)$ is fixed by
the following argument: the energy of the scalar field is given by
$E=\int {\rm d}^3x \sqrt{-g}\rho $ where $\rho =-T^0{}_0$ is the
time--time component of the stress--energy tensor. Requiring that $E$
takes the following (usual) suggestive form in the high-frequency
regime,
\begin{equation}
\label{Energy}
E=\int {\rm d}{\bf k}\frac{\hbar \omega }{2}[
c_{{\bf k}}c_{{\bf k}}^{\dag }+ c_{{\bf k}}^{\dag }c_{{\bf k}}],
\end{equation}
leads to
\begin{equation}
\label{N}
N(k)=\sqrt{\frac{k}{2\hbar }}.
\end{equation}
Therefore, the scalar field operator can be written as:
\begin{eqnarray}
\label{fieldoperator}
\hat{\varphi } (\eta ,{\bf x}) & =& \frac{\sqrt{\hbar }}{a(\eta )}
\frac{1}{(2\pi )^{3/2}} \nonumber \\
& \times & \int \frac{{\rm d}{\bf k}}{\sqrt{2k}}
\biggr[c_{\bf k}(\eta )e^{i{\bf k}\cdot {\bf x}}
+c_{\bf k}^{\dag}(\eta )e^{-i{\bf k}\cdot {\bf x}}\biggl].
\end{eqnarray}
This equation no longer contains arbitrary (or unfixed) factors. The
spirit of this argument is comparable to that of the method
employed in Ref.~\cite{Griden}.
\par
The Hamiltonian can be deduced from the action in a straightforward
manner and reads:
\begin{equation}
\label{Hamiltonian}
H=\frac{\hbar }{2}\int {\rm d}{\bf k}\biggl[k
(c_{\bf k}c_{\bf k}^{\dag }+c_{-{\bf k}}^{\dag }c_{-{\bf k}})
-i\frac{a'}{a}(c_{\bf k}c_{-{\bf k}}
-c_{\bf k}^{\dag }c_{-{\bf k}}^{\dag })\biggr].
\end{equation}
In the above expression, the first term represents the Hamiltonian of
a set of harmonic oscillators, whereas the second term can be viewed
as an interaction term between the perturbations and the
background. This term is present only in a dynamical FLRW Universe since
the coupling function is proportional to the first time derivative of
the scale factor. The time evolution of the field operator is given by
the time evolution of the creation and annihilation operators. It can
be calculated using the Heisenberg equation:
\begin{equation}
\label{Heisenberg}
i\hbar \frac{{\rm d}}{{\rm d}\eta }\hat{\varphi }(\eta ,{\bf x})
=[\hat{\varphi }(\eta ,{\bf x}), H].
\end{equation}
This equation can be solved by a Bogoliubov transformation,
expressed as:
\begin{eqnarray}
\label{Bog1}
c_{\bf k}(\eta ) &=& u_k(\eta )c_{\bf k}(\eta _0)+v_k(\eta )
c_{-{\bf k}}^{\dag }(\eta _0), \\
\label{Bog2}
c_{\bf k}^{\dag }(\eta ) &=& u_k^*(\eta )c_{\bf k}^{\dag }(\eta _0)
+v_k^*(\eta )c_{-{\bf k}}(\eta _0),
\end{eqnarray}
where the functions $u_k(\eta )$ and $v_k(\eta )$ only depend on the
norm of the vector ${\bf k}$. These functions are such that
$|u_k|^2-|v_k|^2=1$, so that the commutation relation between the
creation and annihilation operators is preserved in time. The time
$\eta _0$ must be thought of as the time where the initial conditions are
fixed. Whatever these last ones are, we have $u_k(\eta _0)=1$ and
$v_k(\eta _0)=0$. The differential equations that allow the
determination of $u_k(\eta )$ and $v_k(\eta )$ are:
\begin{equation}
\label{eqsuv}
iu_k'=ku_k+i\frac{a'}{a}v_k^*, \quad iv_k'=kv_k+i\frac{a'}{a}u_k^*.
\end{equation}
If we introduce the Bogoliubov transformation given by
Eqs.~(\ref{Bog1}) and(\ref{Bog2}) in the expression of the field
operator, Eq.~(\ref{fieldoperator}), we obtain:
\begin{eqnarray}
\label{foperatoruv}
\hat{\varphi }(\eta ,{\bf x}) &=& \frac{\sqrt{\hbar }}{a(\eta )}
\frac{1}{(2\pi )^{3/2}}
\int \frac{{\rm d}{\bf k}}{\sqrt{2k}}
\biggl[c_{\bf k}(\eta _0)(u_k+v_k^*)(\eta )e^{i{\bf k}\cdot {\bf x}}
\nonumber \\
& &+c_{\bf k}^{\dag}(\eta _0)(u_k^*+v_k)(\eta )e^{-i{\bf k}\cdot
{\bf x}}\biggr].
\end{eqnarray}
>From Eq.~(\ref{eqsuv}), it is easy to see that the function
$(u_k+v_k^*)(\eta )$ satisfies the same equation as $\chi (\eta ,{\bf
k})$. In the high-frequency regime, the term $a''/a$ becomes
negligible and we have $\lim _{k\rightarrow +\infty }(u_k+v_k^*)(\eta
)=e^{-ik(\eta -\eta _0)}$. This means that, in this regime, the
operator $\hat{\chi }(\eta ,{\bf k})$ is given by:
\begin{eqnarray}
\label{hfchi}
\lim _{k\rightarrow +\infty}\hat{\chi } (\eta ,{\bf k})
= \qquad\qquad\qquad\qquad\qquad\qquad\qquad & & \nonumber \\
\qquad \sqrt{\hbar}
\biggl[c_{\bf k}(\eta _0)
\frac{e^{-ik(\eta -\eta _0)}}{\sqrt{2k}}
+c_{-{\bf k}}^{\dagger}(\eta _0)
\frac{e^{ik(\eta -\eta _0)}}{\sqrt{2k}}\biggr].
\end{eqnarray}
We can now come back to our initial problem, which consists in finding
the correct normalization of the perturbed scalar field and of the
scalar perturbations. We can identify the Fourier component operator
of the perturbed field, $\hat{\varphi }_1(\eta ,{\bf k})$, with
$\hat{\chi }(\eta ,{\bf k})/a(\eta )$, both operators being considered
in the high-frequency regime. Let us emphasize again that this
identification is valid only in this regime. Otherwise, the field
$\hat{\varphi }_1(\eta ,{\bf k})$ does not behave as the free field
$\hat{\varphi }(\eta ,{\bf x})$ and the time dependence of the modes
is no longer given by the function $(u_k+v_k^*)(\eta )$. The
normalization of the perturbed scalar field fixes automatically the
normalization of the scalar perturbations of the metric since they are
linked through Einstein's equations. We only need this link in the
high-frequency regime. It can be expressed as (see
Refs.~\cite{Griden,MS1}):
\begin{equation}
\label{linkmufield}
\lim _{k\rightarrow +\infty }\hat{\mu }(\eta ,{\bf k})=
-\sqrt{2\kappa }a(\eta )\lim _{k\rightarrow +\infty }\hat {\varphi }_1
(\eta ,{\bf k}),
\end{equation}
where $\kappa \equiv 8\pi G$. From the last expression and
Eq.~(\ref{hfchi}), we immediately deduce that:
\begin{eqnarray}
\label{hfmu}
\lim _{k\rightarrow +\infty }\hat {\mu }(\eta ,{\bf k})
= \quad\qquad\qquad\qquad\qquad\qquad\qquad\qquad & & \nonumber \\
-4\sqrt{\pi }l_{\rm Pl}\biggl[c_{\bf k}(\eta _0)
\frac{e^{-ik(\eta -\eta _0)}}{\sqrt{2k}} +c_{-{\bf k}}^{\dagger }(\eta _0)
\frac{e^{ik(\eta -\eta _0)}}{\sqrt{2k}}\biggr],
\end{eqnarray}
where $l_{\rm Pl}=(G \hbar)^{1/2}$ is the Planck length. As announced,
the normalization of the scalar perturbations is now completely
determined. In general, the operator $\hat{\mu }(\eta ,{\bf k})$ will
be given by:
\begin{equation}
\hat{\mu }(\eta ,{\bf k})=-4
\sqrt{\pi }l_{\rm Pl}[c_{\bf k}(\eta _0)\xi _k(\eta )+
c_{-{\bf k}}^{\dag }(\eta _0) \xi _k^*(\eta )]~.
\end{equation}
The function $\xi _k(\eta )$ is the solution of Eq.~(\ref{paraeq})
such that $\lim _{k\rightarrow +\infty } \xi _k =e^{-ik(\eta -\eta
_0)}/\sqrt{2k}$. If we introduce the function $f_k(\eta )$ defined by
\begin{equation}
f_k(\eta )\equiv
-4\sqrt{\pi }[({\cal H}\gamma )/(2k^2)][\xi _k/(a\sqrt{\gamma })']~,
\end{equation}
then the dimensionless Bardeen operator $\hat{\Phi }(\eta ,{\bf x})$
can be written as:
\begin{eqnarray}
\label{operatorphi}
\hat{\Phi }(\eta ,{\bf x}) & = & \frac{l_{\rm Pl}}{(2\pi )^{3/2}} \\
& \times & \int {\rm d}{\bf k}
\biggl[c_{\bf k}(\eta _0)f_k(\eta )e^{i{\bf k}\cdot {\bf x}}
+c_{\bf k}^{\dag }(\eta _0)f_k^*(\eta )e^{-i{\bf k}\cdot {\bf x}}\biggr].
\nonumber
\end{eqnarray}
This equation is the main equation of this section and will be used in
the following.
\par
To conclude this part, it is interesting to consider the previous
equations in the case of power law inflation, i.e. when the scale
factor is given by:
\begin{equation}
\label{defpl}
a(\eta )=l_0|\eta |^{1+\beta }.
\end{equation}
To have inflation, $\beta $ must be such that $\beta +2<0$ (the case
$-2<\beta <-1$ is not considered here because it cannot be realized
with a single scalar field); $\beta =-2$ corresponds to the de Sitter
universe; $l_0$ has dimension of length and, in the de Sitter case, it
is equal to the Hubble radius $l_H\equiv a^2/a'$. Moreover, in a de
Sitter universe, the function $\gamma (\eta )$ turns out to be zero.
We would like to notice
that the assumption of power-law inflation is not as restrictive as it
seems. Indeed, the widely used slow-roll approximation boils down to
power-law inflation, with an effective $\beta $ depending on the slow-roll
parameters, see Refs.~\cite{Lea}. In the case of power law
inflation, analytical exact solutions for the equations of motion of
the perturbations can be found. With the scale factor of
Eq.~(\ref{defpl}), Eq.~(\ref{paraeq}) can be solved in terms of Bessel
functions. Then, the exact solution for the function $\xi _k(\eta )$
is:
\begin{equation}
\xi _k(\eta )=-i(\pi /2)^{1/2}e^{i(k\eta _0-\pi \beta /2)}
(k\eta )^{1/2}H^{(2)}_{\beta +1/2}(k\eta )/\sqrt{2k}~,
\end{equation}
where $H^{(2)}_{\beta +1/2}$ is the second-kind Hankel function of
order $\beta +1/2$. It is straightforward to deduce the corresponding
equation for $f_k(\eta )$:
\begin{equation}
\label{f_kpl}
f_k(\eta ) =
-\pi \sqrt{2}\frac{{\cal H}\sqrt{\gamma }}{k^2}
e^{i(k\eta_0-\frac{\pi \beta }{2})}
\frac{ik}{a}\frac{\sqrt{k\eta}}{\sqrt{2k}}
H^{(2)}_{\beta +3/2}(k\eta ).
\end{equation}
The previous expression for $f_k(\eta)$ guarantees that the field
$\hat{\Phi }(\eta ,{\bf x})$ possesses the correct behaviour in the
high-frequency regime. Roughly speaking, $$\lim _{k\rightarrow +\infty
}f_k(\eta)\sim e^{-ik\eta }/\sqrt{2k},$$ with the correct
amplitude. This should be the case of any field, regardless of the initial
conditions one may choose.
\subsection{Quantum states}
The formulation of quantum field theory used in the previous
subsection was written in the Heisenberg picture. So far, we have only
calculated the time dependence of the Bardeen operator. In order to
describe completely the system, one must in addition specify in which
quantum state the field $\hat{\Phi }(\eta ,{\bf x})$ is placed. As we
already mentioned, it is usually found in the literature that the
initial state of the perturbations is taken to be the vacuum. Here we
address the hypothesis that the perturbations are initially in a non-vacuum
state. Our choice of non-vacuum states is guided by the idea
that one must introduce a scale in the theory. We denote the
corresponding wave number by $k_0$. We examine three different non-vacuum
states.
\par
Let ${\cal D}$ be a domain in the momentum space characterized by the
numbers $k_0$ and $\sigma$, such that $k_0$ is the privileged scale
and $\sigma $ the dispersion around it. Concretely, we take ${\cal D}$
as the space between the spheres of radius $k_0-\sigma $ and
$k_0+\sigma $, i.e. a shell of width $2\sigma $ in $k$-space. The
domain ${\cal D}$ is invariant under rotations and therefore is
compatible with the assumption of isotropy of the Universe. Our first
state is given by:
\begin{eqnarray}
\label{defpsi1}
|\Psi _1(k_0,\sigma ,n)\rangle & \equiv &
\prod _{{\bf k} \in {\cal D}(k_0,\sigma )}
\frac{(c_{\bf k}^{\dagger })^n}{\sqrt{n!}} |0_{\bf k}\rangle
\bigotimes _{{\bf p}\not\in {\cal D}(k_0,\sigma )}|0 _{\bf p}
\rangle , \\
&=& \bigotimes _{{\bf k} \in {\cal D}(k_0,\sigma )}|n_{{\bf k}}\rangle
\bigotimes _{{\bf p} \not\in {\cal D}(k_0,\sigma )}|0_{\bf p}\rangle .
\end{eqnarray}
The state $|n_{{\bf k}}\rangle $ is an $n$-particle state satisfying,
at $\eta =\eta _0$: $c_{{\bf k}}|n_{{\bf k}}\rangle
=\sqrt{n}|(n-1)_{{\bf k}}\rangle $ and $c_{{\bf k}}^{\dag}|n_{{\bf
k}}\rangle =\sqrt{n+1}|(n+1)_{{\bf k}}\rangle $.
\par
More complicated states can be constructed by considering quantum
superpositions of $|\Psi _1(k_0,\sigma ,n)\rangle $. We will consider
the following state:
\begin{equation}
\label{defpsi2}
|\Psi _2(k_0,n)\rangle \equiv \int {\rm d}\sigma g(\sigma )
|\Psi _1(k_0,\sigma ,n)\rangle ,
\end{equation}
where, a priori, $g(\sigma )$ is an arbitrary function of $\sigma
$. It is clear from the definition of the state $|\Psi _1\rangle $
that the transition between the empty and the filled modes is
sharp. Physically, this is probably not very realistic. The function
$g(\sigma )$ will be chosen in order to ``smooth out'' the quantum
state $|\Psi _1\rangle $. Also it should be clear that the writing of
$|\Psi _2\rangle $ in Eq.~(\ref{defpsi2}) is symbolic. An accurate
definition of this state requires to take into account subtleties,
which will be considered when we calculate the spectrum in the next
section.
\par
Finally a third state can be defined according to:
\begin{equation}
\label{defpsi3}
|\Psi _3(k_0)\rangle \equiv \sum _{n=0}^{\infty }h(n)|\Psi
_2(k_0,n)\rangle.
\end{equation}
The function $h(n)$ is arbitrary. As demonstrated below, this state
will allow us to work with an effective number of quanta, which will no
longer be an integer. This state seems to be the most natural
rotational-invariant, smooth, quantum state that privileges a
scale. Typically, the quantum state given in Eq.~(\ref{defpsi3})
depends on $k_0$ and on the free parameters characterizing the
functions $g(\sigma )$ and $h(n)$. Their values will be determined
later by confronting our theoretical predictions versus experimental
and observational data.
\par
Our aim is now to calculate the power spectra of the Bardeen potential
operator in the three states $|\Psi_i\rangle $, $i=1,2,3$.
\subsection{Power spectra}
>From now on, for convenience, we consider that the system is in a box
whose edges have length $L$. As a consequence, the wave vector
possesses discrete components given by $k_i=[(2\pi )/L]m_i$, where
$m_i$ is an integer. The Bardeen operator can be written as:
\begin{eqnarray}
\label{discretephi}
\hat{\Phi }(\eta ,{\bf x})=\frac{l_{\rm Pl}}{L^{3/2}}
\sum _{\bf k}& & [c_{\bf k}(\eta _0)f_k(\eta )e^{i{\bf k}\cdot {\bf x}}
\nonumber \\
& &
+c_{\bf k}^{\dag }(\eta _0)f_k^*(\eta )e^{-i{\bf k}\cdot {\bf x}}].
\end{eqnarray}
We pass from the discrete representation to the continuous one by
sending $L$ to infinity and by applying the rule $1/(2\pi )^{3}\int
{\rm d}{\bf k}\rightarrow 1/L^3\sum _{\bf k}$. It is clear that the
final result does not depend on $L$.
\par
The power spectrum of $\hat{\Phi }(\eta ,{\bf x})$ in the state $|\Psi
\rangle$, denoted by $P_{\Phi }(k;|\Psi \rangle )$, is defined
through the calculation of the two-point correlation function
$K_2(r;|\Psi \rangle )$. In the continuous limit,
\begin{eqnarray}
\label{defK}
K_2(r;|\Psi \rangle )& \equiv &{\langle \Psi |\hat{\Phi }(\eta ,{\bf x})
\hat{\Phi }(\eta ,{\bf x}+{\bf r})|\Psi \rangle} \over
{\langle \Psi |\Psi \rangle}
\nonumber \\
&=& \int _0 ^{\infty }
\frac{{\rm d}k}{k}\frac{\sin kr}{kr}k^3P_{\Phi }(k;|\Psi \rangle).
\end{eqnarray}
In this definition, we have taken into account the fact that the state
$|\Psi \rangle $ is not automatically normalized to 1. The power
spectrum is a time-dependent function but in the long-wavelength limit
this dependence disappears. In order to perform the computation of the
correlation function for the state $|\Psi _1\rangle $, one needs the
following quantities:
\begin{eqnarray}
\label{ccpsi21}
& &\langle \Psi _1(k_0,\sigma ,n)|c_{\bf p}c_{\bf q}|
\Psi _1(k_0,\sigma ,n)\rangle
\nonumber \\
& & =\langle \Psi _1(k_0,\sigma ,n)|c_{\bf p}^{\dag}
c_{\bf q}^{\dag}|\Psi _1(k_0,\sigma ,n)\rangle =0, \\
\label{ccpsi22}
& &\langle \Psi _1(k_0,\sigma ,n)|c_{\bf p}
c_{\bf q}^{\dag}|\Psi _1(k_0,\sigma ,n)
\rangle
=n{\rm \delta }({\bf q}\in {\cal D}){\rm \delta }_{{\bf p}{\bf q}}
+{\rm \delta }_{{\bf p}{\bf q}}, \nonumber \\
& & \\
\label{ccpsi23}
& &\langle \Psi _1(k_0,\sigma ,n)|c_{\bf p}^{\dag }
c_{\bf q}|\Psi _1(k_0,\sigma ,n)
\rangle =n{\rm \delta }({\bf q}\in {\cal D}){\rm \delta }_{{\bf p}{\bf q}}.
\end{eqnarray}
In these formulas, ${\rm \delta }({\bf q}\in {\cal D})$ is a function
that is equal to $1$ if ${\bf q}\in {\cal D}$ and $0$ otherwise.
\par
As a warm up, we calculate the power spectrum for the state $|\Psi
_1\rangle $ with $n=0$, i.e. for the vacuum. Using the previous
equations in the definition of the correlation function,
Eq.~(\ref{defK}), one finds:
\begin{equation}
\label{vac}
K_2(r;|0\rangle )
=\frac{l_{Pl}^2}{L^3}\sum _{\bf k}|f_k|^2e^{-i{\bf k}\cdot {\bf r}},
\end{equation}
which in the continuous limit $L\rightarrow +\infty$ goes to
\begin{equation}
\label{vaccont}
\frac{l_{Pl}^2}{(2\pi )^3}\int d^3\vec{k}|f_k|^2 e^{-i{\bf k}\cdot{\bf r}}.
\end{equation}
After having performed the angular integrations, we recover the power
spectrum $P_{\Phi }(k;|0 \rangle )$:
\begin{equation}
\label{Pvac}
k^3P_{\Phi }(k;|0 \rangle )=\frac{l_{\rm Pl}^2}{2\pi ^2}k^3|f_k|^2.
\end{equation}
Let us turn to the calculation of $K_2(r;|\Psi _1\rangle $. It can be
expressed as:
\begin{equation}
\label{Kpsi1}
K_2(r;|\Psi _1\rangle )
=\frac{l_{Pl}^2}{L^3}\sum _{\bf k }\biggl( |f_k|^2e^{-i{\bf k}\cdot {\bf r}}
[1+2n{\rm \delta }({\bf k}\in {\cal D})]\biggr).
\end{equation}
>From this equation and from the definition of the function ${\rm
\delta }({\bf k}\in {\cal D})$, we deduce the expression of the
power spectrum:
\begin{eqnarray}
\label{Ppsi1}
k^3P_{\Phi }(k;|\Psi _1 \rangle ) &=& \frac{l_{\rm Pl}^2}{2\pi ^2}
k^3|f_k|^2\biggl\{1 +2n\nonumber \\
&\times & [{\rm H}(k-k_0+\sigma )-{\rm H}(k-k_0-\sigma )]\biggr\},
\end{eqnarray}
where $\rm H$ is a Heaviside function. We see that, in addition to the
usual vacuum spectrum, there is a new contribution located around the
wave number $k_0$. This new contribution vanishes if $n=0$, as
expected.
\par
This spectrum is not continuous. As already mentioned, this is not
physically very realistic. It has for origin the very crude definition
of the state $|\Psi _1 \rangle $. We thus turn to the case where the
quantum state is given by $|\Psi _2 \rangle $. This refinement will
allow us to obtain a smooth and physical spectrum.
\par
Since the system is placed in a box, the state $|\Psi _2\rangle $ can
be defined by a discrete sum according to:
\begin{equation}
\label{defpsi2dis}
|\Psi _2(k_0,n)\rangle \equiv
\sum _{i=0}^{N} g_i|\Psi _1(k_0,\sigma _i,n)\rangle ,
\end{equation}
where $g_i$ and $\sigma _i$, $i=0, \dots ,N$, are just series of
numbers. We choose the $\sigma _i$'s such that:
\begin{equation}
\label{orthocondition}
\langle \Psi _1(k_0,\sigma _i,n)|\Psi _1(k_0,\sigma _j,n)\rangle
=\delta _{ij}.
\end{equation}
This is satisfied if the number of modes in the domains ${\cal
D}(k_0,\sigma _i)$ and ${\cal D}(k_0,\sigma _j)$, ${\cal N}({\cal
D}_i)$ and ${\cal N}({\cal D}_j)$ respectively, are such that: ${\cal
N}({\cal D}_i)-{\cal N}({\cal D}_j)\ge 1$. This condition boils down
to $\sigma _i -\sigma _j \ge \pi ^2/ [L^3(k_0^2+\sigma _i^2)]$, where
we have assumed that $\sigma _i- \sigma _j\ll 1$. Therefore, we can
always find a value of $L$ such that the condition be fulfilled. Then,
the calculation of $\langle \Psi _2|\hat{\Phi }(\eta ,{\bf
x})\hat{\Phi }(\eta ,{\bf x}+{\bf r})|\Psi _2\rangle $ can be
performed. The result reads:
\begin{eqnarray}
\label{Kpsi2}
& &\langle \Psi _2|\hat{\Phi }(\eta ,{\bf x})
\hat{\Phi }(\eta ,{\bf x}+{\bf r})|\Psi _2\rangle
=\frac{l_{Pl}^2}{L^3}\sum _{\bf k }
|f_k|^2e^{-i{\bf k}\cdot {\bf r}} \nonumber \\
& &\times \biggl\{\biggl[\sum _{i=0}^{N}|g_i|^2\biggr]
+2n\biggr[\sum _{i=0}^{N}|g_i|^2
{\rm \delta }\biggl({\bf k}\in {\cal D}(k_0,\sigma _i)\biggr)\biggr]\biggr\}.
\end{eqnarray}
Our aim is to calculate $G(k) \equiv \sum _{i=0}^N|g_i|^2{\rm \delta
}({\bf k}\in {\cal D}_i)$ [for convenience ${\cal D}(k_0,\sigma _i)$
is denoted by ${\cal D}_i$]. By symmetry, $G(k_0-k')=G(k_0+k')$, so
that we will consider the case $k \equiv k_0 + k'$, $k'\ge 0$. In this
sum $k'$ is fixed. As a consequence, there exists an integer $i_0$
such that if $i < i_0$, ${\rm \delta }({\bf k}\in {\cal D}_i)=0$ and
if $i\ge i_0$, then ${\rm \delta }({\bf k}\in {\cal D}_i)=1$, or,
equivalenty, $\sigma_{i_0} < k'\le \sigma_{i_0+1}$. This means that
the sum $\sum _{i=0}^N|g_i|^2{\rm \delta }({\bf k}\in {\cal D}_i)$ is
in fact equal to $\sum _{i=i_0}^N|g_i|^2$. We choose the $\sigma_i$'s
and the coefficients $g_i$ according to:
\begin{eqnarray}
\label{defgi}
\sigma_i \equiv i\frac{X_{\rm max}}{N}\quad,\quad
|g_i|^2 \equiv - \frac{X_{\rm max}}{N} \times
\left.\frac{{\rm d}F}{{\rm d}x} \right|_{x=\sigma_i},
\end{eqnarray}
where $F$ is any decreasing function such that $F(X_{\rm
max})=0$. Then we have
\begin{equation}
\label{defgi2}
\sum_{i=0}^N|g_i|^2{\rm \delta }({\bf k}\in {\cal D}_i)
= - \frac{X_{\rm max}}{N}
\sum _{i=i_0}^N F'\left(i\frac{X_{\rm max}}{N}\right).
\end{equation}
The last step is to send $N$ to infinity. This means that we consider
a continuous series of intervals ${\cal D}_i$. We obtain:
\begin{eqnarray}
\label{Riemansum}
\lim _{N\rightarrow +\infty }
\sum _{i=0}^N|g_i|^2{\rm \delta }({\bf k}\in {\cal D}_i)
& = & - \int_{k'}^{X_{\rm max}} F'(x){\rm d}x \nonumber \\
& = & F(k'),
\end{eqnarray}
by definition of the Riemann integral. In the same manner, $\sum
_{i=0}^N|g_i|^2=F(0)$. In what follows, we take
\begin{equation}
\label{defF}
G(k) = F(k')\equiv e^{-\frac{(k-k_0)^2}{\Sigma ^2}},
\end{equation}
where $\Sigma $ is a free parameter. This function does not exactly satisfy
the assumptions made previously, but it is easy to show that
the final result is free of these limitations and is in fact valid for
any function $F$. It is clear that other functions are possible, but
only the approximate shape of the distribution is important and the
Gaussian is the prototype of the function we have in mind. The
spectrum is obtained after having introduced the previous result in
Eq.~(\ref{Kpsi2}) and having taken the limit $L\rightarrow
+\infty$. We obtain:
\begin{equation}
\label{Ppsi2}
k^3P_{\Phi }(k;|\Psi _2 \rangle )=\frac{l_{\rm Pl}^2}{2\pi ^2}
k^3|f_k|^2\biggl(1 +2ne^{-\frac{(k-k_0)^2}{\Sigma ^2}}\biggr).
\end{equation}
In this equation, it is clear that $n$ is an integer. We now show that
this condition can be relaxed if the system is placed in the state
$|\Psi _3 \rangle $.
\par
To calculate the spectrum for this state, it is sufficient to notice
that $\langle \Psi _2(k_0,n)|\Psi _2(k_0,m)\rangle =\delta
_{mn}$. Using this formula, straightforward calculations lead to:
\begin{equation}
\label{Ppsi3}
k^3P_{\Phi }(k;|\Psi _3 \rangle )=\frac{l_{\rm Pl}^2}{2\pi ^2}
k^3|f_k|^2\biggl(1 +2n_{\rm eff}e^{-\frac{(k-k_0)^2}{\Sigma ^2}}\biggr),
\end{equation}
where the effective number of quanta, $n_{\rm eff}$, is given by:
\begin{equation}
\label{defneff}
n_{\rm eff}=\frac{\sum _{n=0}^{\infty }n|h(n)|^2}
{\sum _{n=0}^{\infty }|h(n)|^2}.
\label{neff}
\end{equation}
An attractive choice for the function $h(n)$ is obviously $h(n)\equiv
e^{-\beta n}$ [this $\beta$ has of course nothing to do with the
$\beta$ defined in Eq.~(\ref{defpl})]. In this case, $n_{\rm eff}$ is
given by:
\begin{equation}
\label{thermalneff}
n_{\rm eff}=\frac{e^{-2\beta }}{1-e^{-2\beta }}.
\end{equation}
The spectra of Eqs.~(\ref{Ppsi2}) and (\ref{Ppsi3}) are the main
results of this section. Clearly, they possess a peak around the scale
$k_0$. The position of the peak is controlled by the value of $k_0$,
its width by $\Sigma$ and its height by $n$ or $n_{\rm eff}$ (in fact
by $\beta $ in the last case).
\par
We will need the primordial spectrum only for large wavelengths. In the
case of power law inflation, everything can be calculated exactly. In
this limit, we have
\begin{equation}
\label{An0}
k^3P_{\Phi}(k;|0\rangle )=A_{\rm S}k^{n_{\rm S}-1}~,
\end{equation}
with
\begin{equation}
\label{An1}
A_{\rm S}=\frac{l_{\rm Pl}^2}{l_0^2}\frac{\gamma (1+\beta
)^2}{2^{2\beta +4}
\cos ^2(\beta \pi )\Gamma ^2(\beta +5/2)}, \quad n_{\rm S}=2\beta +5.
\end{equation}
The above expression is strictly speaking not applicable in the case of a
de Sitter universe, since then there are no scalar metric perturbations and
the function $\gamma(\eta)$ turns out to be zero. The generation of density
perturbations is only possible after the transition from the exponential
inflationary era to the radiation-dominated Universe. If during inflation the
Universe was very close to the de Sitter space-time, then the spectrum of
density perturbations today is the so-called Harrison-Zel'dovich spectrum
($n_{\rm S}=1$). All expressions derived in this section are still valid for
$\beta {\raise 0.4ex\hbox{$<$}\kern -0.8em\lower 0.62
ex\hbox{$\sim$}} -2$.
The initial power spectrum in the case where the Bardeen
operator is placed in the state $|\Psi _2\rangle $ can be written as:
\begin{equation}
\label{Ppsi3lwl}
k^3P_{\Phi}(k;|\Psi _2\rangle )=
A_{\rm S}k^{n_{\rm S}-1}
\left(1+2n e^{-\frac{(k-k_0)^2}{\Sigma ^2}}\right).
\end{equation}
If the state is $|\Psi _3\rangle$, we just have to replace the integer
$n$ with the real number $n_{\rm eff}$. Let us note that if, instead
of considering intervals of the form $[k_0-\sigma ,k_0+\sigma ]$, one
considers intervals such as $]0,k_0+\sigma ]$ or $[k_0-\sigma ,\infty[$,
which still privilege a scale, it is possible to build step-like
spectra, the step being located at the scale $k_0$.
\par
Recently in the literature, BSI spectra have
also been studied, see Ref.~\cite{bsi}. In these articles, the
privileged scale arises as a privileged energy in the inflaton
potential (more precisely, a discontinuity, or a rapid variation, in the
inflaton potential if first derivatives are present). We would like to
stress that, in our case, the different physical origin of the
privileged scale would in principle allow us to distinguish the different
models. Indeed, in Ref.~\cite{bsi}, the fluctuations are Gaussian. In
the case studied here, the three-point correlation function still
vanishes
\begin{equation}
\label{3points}
\left\langle{{\rm \delta} T\over T}({\bf e}_1)
{{\rm \delta} T\over T}({\bf e}_2)
{{\rm \delta} T\over T}({\bf e}_3)\right\rangle = 0 \;~,
\end{equation}
but the four-point correlation function no longer satisfies the
relation
\begin{equation}
\label{4points}
\left\langle\biggl({{\rm \delta} T\over T}({\bf e})\biggr)^4\right\rangle
= 3 \Biggl[\left\langle\biggl({{\rm \delta} T\over T}({\bf
e})\biggr)^2\right\rangle\Biggr] ^2,
\end{equation}
which is typical of Gaussian statistics. The reason for this is
clear. The ground-state wave function of an harmonic oscillator is a
Gaussian and, as a consequence, the CMBR correlation functions for the
vacuum exhibit Gaussian properties. On the other hand, the wave
function of a state with a non-vanishing number of quanta is no longer
a Gaussian and, correspondingly, the correlation functions deviate from
Gaussianity. Therefore, a measure of the four-point correlation
function (as well as any higher-order even-point correlation
function) would permit to distinguish between the class of models
presented here and the models of Ref.~\cite{bsi}. If it turns out that
the type of non-Gaussianity apparently detected recently
\cite{non-gaus1,non-gaus2,non-gaus3} is really present in the CMBR
map, then these two classes of BSI models (as well as standard
inflation) are ruled out, because they both predict a vanishing
three-point correlation function. But if it turns out that some
non-Gaussianity is present in the CMBR at the level of the four-point
correlation function then the models presented
here could account for this.
\section{Comparison with observations}
The aim of this section is to confront the power spectra given by
Eq.~(\ref{Ppsi3lwl}) with observations. We will not use any accurate
statistical methods to find the best values of the free parameters
$k_0$, $\Sigma $ and $n$/$n_{\rm eff}$, because we just want to obtain
crude constraints. For this purpose we will use observations of the
CMBR anisotropies and of the matter power spectrum.
\par
We choose to work with the following cosmological parameters: the
Hubble parameter is $h=0.5$, the baryonic matter-density parameter is
$\Omega _b=0.05$, the density parameter $\Omega_0 \equiv
\Omega _c+\Omega _b+\Omega_\Lambda$ is equal to 1 ($\Omega_c$
and $\Omega_\Lambda$ are respectively the CDM and the $\Lambda$-density
parameters), there is no significant reionization and the spectral
index is $n_{\rm S}=1$, when
there is no contribution from the gravitational waves.
Let us emphasize again that in this case, since it corresponds to a de
Sitter phase, the Eq.~(\ref{An1}) giving the normalization
of the spectrum is strictly speaking not applicable.
However, all expressions derived in the previous section can be applied for
$n_{\rm S} {\raise 0.4ex\hbox{$<$}\kern -0.8em\lower 0.62
ex\hbox{$\sim$}} 1$. We will later discuss the case of small deviations
from a scale-invariant (Harrison-Zel'dovich) spectrum, including the
contribution of gravitational waves in the CMBR anisotropies.
We will consider
two different values for
the cosmological constant density parameter $\Omega _{\Lambda }\equiv
\Lambda/(3H_0^2)$ and the sum of baryon-matter density parameter and
CDM density parameter $\Omega _m\equiv \Omega _b+\Omega
_c$, that is $\Omega_\Lambda=0$, $\Omega_c = 0.95$ (hereafter denoted
SCDM), and $\Omega_\Lambda=0.6$, $\Omega_c = 0.35$ (hereafter denoted
$\Lambda$CDM).
We point out that we have not assumed any biasing in the galaxy distribution
with respect to the underlying mass fluctuations (the bias parameter is equal to 1).
\par
The spectrum must be normalized, i.e. the value of $A_{\rm S}$ must be
determined. For this purpose, we use the value of $Q_{\rm
rms-PS}=T_0(5C_2/4\pi )^{1/2} \sim 18 ~{\rm \mu K}$ ($T_0=2.7 ~{\rm K}$)
measured by the COBE satellite. We use the value $Q_{\rm rms-PS}\sim
18 ~{\rm \mu K}$ because we have assumed that the spectrum is
scale-invariant. In the large-wavelength approximation, we have ${\rm \delta
}T/T\sim (1/3)\Phi $. In addition the transfer function for the Bardeen
potential can be taken equal to 1 (with an appropriate
normalization). As a consequence the
multipole can be written as:
\begin{eqnarray}
\label{cl}
C_\ell & = & \frac{4\pi }{9}\int _0^{+\infty }\frac{{\rm d}k}{k}
\biggl[j_\ell[k(\eta_0-\eta_{\rm LSS})]^2 \nonumber \\
& & \quad \times
A_{\rm S}(n_{\rm S})k^{n_{\rm S}-1}
\left(1+2ne^{-\frac{(k-k_0)^2}{\Sigma ^2}}\right)
\biggr],
\end{eqnarray}
where $j_\ell$ is a spherical Bessel function of order $\ell$, and
$\eta_0$ and $\eta_{\rm LSS}$ denote respectively the conformal times
now and at the last scattering surface. Let us remark that the $A_{\rm
S}$ in the last expression is not exactly the $A_{\rm S}$ in
Eqs.~(\ref{An0}) and (\ref{An1}). Since the difference is not important
for our purpose, we have kept the same notation. The previous
expression can be evaluated explicitly. For the quadrupole, the result
reads:
\begin{eqnarray}
\label{quadrupole}
& & C_2 =\frac{4 \pi}{9}
A_{\rm S}(n_{\rm S})(\eta_0-\eta_{\rm LSS})^{1-n_{\rm S}}
\nonumber \\ & & \times \biggl(\frac{\pi}{2^{4-n_{\rm S}}}
\frac{\Gamma [3-n_{\rm S}]\Gamma [2+(n_{\rm S}-1)/2]}
{\Gamma ^2[(4-n_{\rm S})/2]\Gamma [4-(n_{\rm S}-1)/2]}
+2n I \biggr) ,
\end{eqnarray}
where
\begin{equation}
\label{I}
I\equiv \int _0^{+\infty }
\frac{{\rm d}u}{u^{2-n_s}}
[j_2(u)]^2
e^{-\frac{(u-u_0)^2}{U^2}},
\end{equation}
with $u_0 \equiv k_0(\eta_0-\eta_{\rm LSS})$ and $U \equiv
\Sigma(\eta_0-\eta_{\rm LSS})$. In what follows, we take
$\eta _0-\eta _{LSS}=1$. The integral will be evaluated numerically
for different values of the free parameters. We just have to
specialize the last equation to a scale-invariant spectrum to obtain
the following value for $A_{\rm S}$:
\begin{equation}
\label{A_S}
A_{\rm S}=\frac{108}{5T_0^2}Q^2_{\rm rms-PS}\frac{1}{1+24 n I}=
\frac{9.4\cdot 10^{-10}}{1+24 n I}.
\end{equation}
In terms of the band power ${\rm \delta }T_\ell$ defined by ${\rm
\delta }T_\ell\equiv T_0[\ell(\ell+1)C_\ell/2\pi]^{1/2}$, we find
${\rm \delta }T_2=\sqrt{12/5}Q_{\rm rms-PS}=27.9 ~{\rm \mu K}$.
\par
We must choose the three parameters $k_0$, $n$ and $\Sigma
$. Recently, it has been emphasized by many authors \cite{einasto}
that the power
spectrum seems to contain large amplitude features at the scale
$l_{\rm C}\approx 100 ~h^{-1}~{\rm Mpc}$, which corresponds to a wave number
equal to $0.062 ~h~{\rm Mpc}^{-1}$. No other value for a privileged scale has
been detected so far, and therefore any other choice would either lie
in an unobservable range, or be in conflict with the data available at
present. Consequently, we choose:
\begin{equation}
\label{position}
k_0=0.062 ~h~ {\rm Mpc}^{-1}=0.031 ~{\rm Mpc}^{-1},
\end{equation}
with our value of the Hubble constant. Let us turn to the choice of
the variance $\Sigma$. We have seen that the simplest non-vacuum
initial states can lead to a power spectrum with either a bump or a
step. In this article, we will restrict ourselves to the study of the
bump case. Step-like spectra have already been studied in
Ref.~\cite{bsi} and our conclusions would be similar. Therefore we
will consider (rather arbitrarily, but the conclusion does not depend on
the exact value of $\Sigma$, as long as it is not too large):
\begin{equation}
\label{variance}
\Sigma =0.3k_0=0.0186 ~h~ {\rm Mpc}^{-1} .
\end{equation}
>From now on, we will always take these two values for $k_0$ and
$\Sigma$ in any of the plots shown. In Fig.~\ref{ps_1} we display the
initial power spectrum for a few values of $n_{\rm eff}$. The
difference between $n_{\rm eff}=0$ and $n_{\rm eff}\neq 0$ is obvious.
\begin{figure}
\begin{center}
\leavevmode
\input{ps_in}
\end{center}
\caption{Initial power spectrum for $n_{\rm eff}$ ranging from 0 to
2 with steps of 0.5. Vertical units are arbitrary.}
\label{ps_1}
\end{figure}
In the case considered here, the integral $I$ is equal to: $I(\Sigma
=0.3k_0)\approx 1.3 \cdot 10^{-6}$. It is completely negligible and
will be taken equal to zero. This arises from the fact that the
quadrupole is mainly fed by very large wavelengths (of the order of today's
Hubble radius), whereas the bump occurs at much smaller wavelengths
(of the order of the Hubble radius at the time of decoupling). Thus, the
calculation of the quadrupole, and therefore the normalization, is not
modified by the presence of the bump.
\par
Let us discuss the matter power spectrum. The power spectrum can
either be obtained by the Boltzmann code developed by one of us
(A.~R.) or by means of analytical fits. In this case,
the baryons power spectrum is given by:
\begin{equation}
\label{psmatter}
\frac{{\rm \delta }\rho _{\rm b}}{\rho _{\rm b}}\equiv |\delta (k)|^2=AT^2(k)
\frac{g^2(\Omega _0)}{g^2(\Omega_m)}k\biggl[1+2ne^{-\frac{(k-k_0)^2}
{\Sigma ^2}}\biggr],
\end{equation}
where the different terms in this equation are explained below; $T(k)$
is the transfer function, which can be approximated by the following
numerical fit~\cite{bbks}:
\begin{eqnarray}
\label{transfert}
T(k) & = & \frac{\ln (1+2.34 q)}{2.34 q}[1+3.89q \nonumber \\
& & +(16.1q)^2+(5.46q)^3+(6.71q)^4]^{-1/4},
\end{eqnarray}
with $q\equiv k/[(h\Gamma ){\rm Mpc}^{-1}]$ where $\Gamma $ is the
so-called shape parameter, which can be written as~\cite{sugiyama}:
\begin{equation}
\label{shapepara}
\Gamma \equiv \Omega _mhe^{-\Omega _b-\frac{\Omega _b}{\Omega _m}}.
\end{equation}
The function $g(\Omega)$ takes into account the modification induced
in the power spectrum by the presence of a cosmological constant. Its
expression can be written as~\cite{primack}:
\begin{equation}
\label{functiong}
g(\Omega )\equiv \frac{5\Omega }{2}\biggl[\Omega ^{4/7}-\Omega _{\Lambda }+
\biggl(1+\frac{\Omega }{2}\biggr)\biggl(1+\frac{\Omega _{\Lambda }}{70}\biggr)
\biggr]^{-1}.
\end{equation}
Finally the coefficient $A$ is the normalization. We normalize
the spectrum to COBE data. This leads to the following value for $A$:
\begin{eqnarray}
\label{A}
A &=& (2l_H)^4\frac{6\pi ^2}{5}
\frac{Q_{\rm rms-PS}^2}{T_0^2}\frac{1}{1+24 n I} \\
&=& \frac{6.82\cdot 10^5}{1+24 nI}h^{-4}{\rm Mpc}^4,
\end{eqnarray}
where the Hubble radius, $l_H$, is equal to $3000h^{-1}{\rm Mpc}$.
\par
We plot the multipole moments and the power spectrum for different
values of $n$ and/or $n_{\rm eff}$. The $C_\ell$'s are obtained from
the Boltzmann code previously used for the power spectrum. In all
figures for the $C_\ell$'s, we represent the COBE data~\cite{FIRS} by
diamonds, the Saskatoon data~\cite{Saskatoon} by squares, and the
CAT~\cite{CAT} data by crosses. (For clarity we have not displayed all
CMBR data on the figures.) In all figures for the power spectra, we
represent the APM data~\cite{APM} by diamonds, the velocities field
measurements~\cite{vel} by squares, and the data given by Einasto {\it
et al.}~\cite{einasto} by crosses.
\subsection{Scalar modes only}
We first display the CMBR anisotropies in the SCDM model
(Fig.~\ref{cl_1}).
\begin{figure}
\begin{center}
\leavevmode
\input{cl_s1_fl}
\end{center}
\caption{Multipole moments for the SCDM model with $n_{\rm eff}$ (and
$n$ if it is integer) ranging from 0 to 2 with step of 0.5 (from the
bottom to the top). Diamonds represent COBE data, squares the
Saskatoon data, and crosses the CAT data.}
\label{cl_1}
\end{figure}
In the case were $\Lambda=0$, Saskatoon data are compatible with the
case $n_{\rm eff}=1$ (third curve).
We note that the position of the first Doppler peak is no longer
around $\ell \approx 220$. Usually, its position is determined by the
angular size of the Hubble radius at recombination. In our case, we
must superimpose the bump present in the initial spectrum, the
position of which is not at $\ell \approx 220$ but rather at the
angular scale sustained by the built-in scale. As a consequence, the
resulting peak is shifted towards higher values of $\ell$ for the
values of the parameters considered here ($\ell \approx 260$). In
addition, it could be difficult to distinguish the effect due to the
primordial bump from the one coming from a variation of the
cosmological parameters, thus increasing the degeneracy among the free
parameters of the model. Let us note, however, that the bump in the
initial power spectrum, should be easier to detect in the matter power
spectrum since it is a more slowly varying function, as shown in
Fig.~\ref{ps_2}.
\begin{figure}
\begin{center}
\leavevmode
\input{ps_s1_fl}
\end{center}
\caption{Power spectrum for the SCDM model, with $n_{\rm eff}$ ranging
from 0 to 2 with step of 0.5 (from the bottom to the top). Diamonds
represent the APM data, squares the velocities field
measurements, and crosses the data by Einasto {\it et
al.} }
\label{ps_2}
\end{figure}
A higher value of $n_{\rm eff}$ (2 rather than 1) seems to be needed
to explain the data of Einasto {\it et al.}, but different cosmological
parameters might
lead to a better agreement between CMBR and matter power spectrum
data.
\par
We now display the CMBR (Fig.~\ref{cl_2}) and matter power spectrum
(Fig.~\ref{ps_3}) of the $\Lambda$CDM model.
\begin{figure}
\begin{center}
\leavevmode
\input{cl_s1_lm}
\end{center}
\caption{Same as Fig.~\ref{cl_1}, but for the $\Lambda$CDM model, with
$\Omega_\Lambda = 0.6$.}
\label{cl_2}
\end{figure}
\begin{figure}
\begin{center}
\leavevmode
\input{ps_s1_lm}
\end{center}
\caption{Same as Fig.~\ref{ps_2}, but for the $\Lambda$CDM model.}
\label{ps_3}
\end{figure}
When the cosmological constant $\Omega_\Lambda = 0.6$, the early Integrated
Sachs--Wolfe effect already boosts the $\ell\simeq 200-300$ scale
sufficiently \cite{Hu_1}: at $n_{\rm eff}=1$ this effect already puts too
much power on these scales. A
different value for $k_0$ and $\Sigma$ might also be needed
to remain compatible with the CAT data. For the matter power
spectrum, the same conclusion as for the SCDM model holds, that is a
higher value of $n_{\rm eff}$ is preferred (around 2 or 3).
\par
As a conclusion of this rapid analysis, we stress that our model is
much more constrained if one imposes $n_{\rm eff}$ to be an integer
instead of a real number. Moreover, our model tends to favour a moderate
value of $n_{\rm eff}$ as well as a low value of the cosmological
constant if the data of Einasto {\it et al.} are confirmed, or a
low value of $n_{\rm eff}$ and a high value of the cosmological constant
(i.e.~the currently popular cosmological model, with vacuum initial state) in
the other case. It is easy to notice from Eq.~(\ref{neff}),
that since $n_{\rm eff}$ is quite small, $h(n)$ is peaked around small
values of $n$, and therefore the allowed window for the effective number of
quanta is constrained to be around small values. In conclusion,
the initial state found is not too far from the vacuum.
\subsection{Scalar and tensor modes}
One should also consider the contribution of the gravitational waves
in the CMBR anisotropies. The data currently available are in fact the
sum of the scalar plus the tensor contributions to the CMBR
anisotropies. We recall that since there are two modes of polarization
for the CMBR photons and that one of them is only generated by
gravitational waves, it is in principle possible to distinguish
between the scalar and tensor contributions to the CMBR anisotropies,
see~\cite{Hu_2}. In what follows, we consider some standard
inflationary predictions for gravitational waves: we take $n_{\rm S} =
0.9$, $n_{\rm T} = n_{\rm S}-1 \approx -0.1$ (the last equation being
rigorous in the case of power-law inflation only), and the ratio of
scalar to tensor amplitude $C_2^{\rm T} / C_2^{\rm S}
\approx - 7 n_{\rm T}$. In Fig.~\ref{cl_3}, we decompose CMBR
anisotropies, showing the contributions from scalar and
tensor modes separately.
\begin{figure}
\begin{center}
\leavevmode
\input{cl_st_de}
\end{center}
\caption{CMBR anisotropies decomposition, showing scalar (dotted
line) and tensor (dashed line) contributions. The total contribution
is given by the solid line.}
\label{cl_3}
\end{figure}
In any model, the gravitational waves contribution can be important only
for multipoles $\ell ~\lower2pt\hbox{$\buildrel{<}\over{\sim}$} ~ 100$, while it is negligible at smaller
angular scales (roughly speaking, the gravitational waves contribution
is two orders of magnitude smaller at $\ell \approx 300$ than at the
quadrupole). The effect of gravitational waves is therefore to boost
power on large angular scales (or, equivalently, to lower the height of
the acoustic peaks with respect to the height of the low $\ell$
plateau). The fact that one observes an excess of power on small
angular scales (with Saskatoon data), favours a low contribution from
gravitational waves (which is in agreement with most inflationary
models). In our model, the possibility to have a bump in the initial
power spectrum enables us to boost the height of the acoustic peaks,
and therefore to have some non-negligible contribution from
gravitational waves: normalizing at COBE data, one imposes the value
of $A_{\rm S} + A_{\rm T}$ instead of $A_{\rm S}$. As a result, the
scalar perturbations amplitude $A_{\rm S}$ is smaller. Since the first
acoustic peak depends only on scalar perturbations, we must keep the
same value as before for the product $A_{\rm S}(1+2n\exp[-(k_0-k_{\rm
peak})^2/\Sigma^2])$, which permits a higher value of $n_{\rm eff}$
($k_{\rm peak}$ is the characteric wave number of the first Doppler
peak).
\par
In Figs.~\ref{cl_4} and \ref{ps_4} we show the CMBR anisotropies and the
matter power spectrum for the SCDM model, including both scalar and
tensor contributions.
\begin{figure}
\begin{center}
\leavevmode
\input{cl_st_fl}
\end{center}
\caption{CMBR anisotropies for the SCDM model, with $n_{\rm eff}$
ranging from 0 to 4 with a step of 1 (from the bottom to the
top). Both scalar and tensor contributions are included. Diamonds
represent COBE data, squares the Saskatoon data, and crosses
the CAT data.}
\label{cl_4}
\end{figure}
\begin{figure}
\begin{center}
\leavevmode
\input{ps_st_fl}
\end{center}
\caption{Power spectrum for the SCDM model, with $n_{\rm eff}$ ranging from
0 to 4 with step of 1. Diamonds represent the APM data, squares the
velocities field measurements, and crosses the data given by
Einasto {{\it et al.}}}
\label{ps_4}
\end{figure}
Comparing these figures with Figs.~\ref{cl_1} and\ref{ps_2}, it can be
concluded that if both scalar and tensor modes are included in the
calculation of the multipole moments $C_\ell$'s, then a higher number
of quanta ($\simeq 4$) is required as expected.
\par
Finally, in Figs.~\ref{cl_5} and \ref{ps_5} we show the CMBR anisotropies
and the matter power spectrum for the $\Lambda$CDM model including
both scalar and tensor contributions.
\begin{figure}
\begin{center}
\leavevmode
\input{cl_st_lm}
\end{center}
\caption{Same as Fig. \ref{cl_4}, but for the $\Lambda$CDM model and
with $n_{\rm eff}$ ranging from 0 to 4 with a step of 1 (from the
bottom to the top).}
\label{cl_5}
\end{figure}
\begin{figure}
\begin{center}
\leavevmode
\input{ps_st_lm}
\end{center}
\caption{Same as Fig. \ref{ps_4}, but for the $\Lambda$CDM model and
with $n_{\rm eff}$ ranging from 0 to 4 with a step of 1 (from the
bottom to the top).}
\label{ps_5}
\end{figure}
The same conclusions as for the SCDM model hold, but we note again
that, as for the case without gravitational waves, matter power
spectrum data favour a higher value of $n_{\rm eff}$ than CMBR
anisotropies data.
\par
We emphasize that when the gravitational waves contribution is not
negligible, the standard case $n_{\rm eff} = 0$ is excluded and that
extra power in the initial state is necessary.
\section{Conclusions}
In this paper we address the question of whether non-vacuum initial
states for cosmological perturbations are allowed, or whether they are
ruled out on the basis of present experimental and observational
data.
\par
The choice of the initial quantum state in which the quantum fields are
placed should be made on the basis of full quantum gravity. Since this
theory is at present unknown, we believe, as we discussed in the
Introduction, that it is worth studying non-vacuum initial states for
cosmological perturbations. Our choice of a non-vacuum initial state
is guided by the idea that the initial state could have a built-in
characteristic scale. We examined three
different non-vacuum states,
which are compatible with the assumption of isotropy of the
Universe. Of particular interest is our choice of state $|\Psi_3
\rangle $, which seems to be the most natural rotational-invariant
smooth quantum state, which privileges a scale. We calculated the power
spectra of the Bardeen potential for these three states and compared
their theoretical predictions with current experimental and
observational data, namely the CMBR anisotropy measurements and the
redshift surveys of the distribution of galaxies. With our choice of
initial states, the power spectra of the Bardeen potential possess a
peak, around the wave number, that corresponds to the built-in
characteristic scale of our model. The height of the peak is
controlled by the number of quanta $n$ of the initial state and its
width by another free parameter of our model. If the initial state
is a quantum superposition then the height of the peak is
controlled by the number $n_{\rm eff}$, which does not need
to be an integer.
\par
The angular power spectrum of CMBR anisotropies for a model with
vanishing cosmological constant, tells us that the
characteristics of the first acoustic peak, as revealed by the
Saskatoon experiment, are compatible with the case $n_{\rm
eff}=1$. In the presence of a cosmological constant, CMBR anisotropy
measurements are in agreement with $n_{\rm eff}=0$ or $n_{\rm
eff}=1$, depending on the value of the cosmological parameters. The
observational data for the matter power spectra, as given by Einasto
{\it et al.}, favour higher values of the number of $n_{\rm eff}$ (2
or 3), whatever the value of the cosmological constant.
\par
The most realistic case is the one for which the sum of scalar and
tensor modes contributions is included. Considering standard
inflationary predictions for gravitational waves, we find that CMBR
anisotropies measurements require a higher value of $n_{\rm eff}$ (3
or 4) for both types of models, with and without a cosmological
constant, than in the case of an absence of tensor modes
contribution. This is in agreement with the matter power spectra. The
analysis of the redshift surveys by Einasto {\it et al.} leads
to matter power spectra that favour higher values of $n_{\rm eff}$,
once tensor contributions are also included. The interpretation of
these results for the states $|\Psi _2\rangle $ and $|\Psi _3\rangle $
leads to the conclusion that since
$n$ and $n_{\rm eff}$ cannot be higher than a few, these states must be
close to the vacuum.
\par
In conclusion, if the initial state of the cosmological perturbations is
not the vacuum but, instead, has a built-in characteristic scale, then
generic predictions of the model are: a high amplitude of the first
acoustic peak, a non-trivial feature in the matter power spectrum, and
deviations from Gaussianity in the CMBR map. It is too early to say
whether the results of the Saskatoon experiment (see also
Ref. \cite{PythonV}), as well as the analysis performed recently by
Einasto {\it et al.}, are first steps in this direction. More data are
needed and future experiments will be important in determining whether
the class of models proposed here provides an explanation which allows
a better description of the observations than the standard paradigm of
slow roll inflation plus cold dark matter.
\vspace{0.5cm}
\noindent
\acknowledgements
It is a pleasure to thank Robert Brandenberger for useful exchanges of
comments. Discussions with Nathalie Deruelle, Ruth Durrer, Alejandro Gangui
and David Langlois are also acknowledged.
We would also like to thank Volker M\"uller, who provided us with the cluster
data.
|
\section{\@startsection {section}{1}{\setbox0\hbox{+}\hbox to \wd0{\hss0\hss}@}{+3.0ex plus +1ex minus
+.2ex}{2.3ex plus .2ex}{\normalsize\bf}}
\def\subsection{\@startsection{subsection}{2}{\setbox0\hbox{+}\hbox to \wd0{\hss0\hss}@}{+2.5ex plus +1ex
minus +.2ex}{1.5ex plus .2ex}{\normalsize\bf}}
\def\subsubsection{\@startsection{subsubsection}{3}{\setbox0\hbox{+}\hbox to \wd0{\hss0\hss}@}{+3.25ex plus
+1ex minus +.2ex}{1.5ex plus .2ex}{\normalsize\bf}}
\def\Alph{section}}.\arabic{equation}{\Alph{section}}.\arabic{equation}}
\@addtoreset{equation}{section}
\def |
\section{Introduction}
\setcounter{theorem}{0}
In this section we present several results concerning intresections of
filters on $\omega$.
The following classical result is a starting point for all subsequent
theorems.
\begin{theorem}[Sierpinski]\label{sierpinski}
Suppose that ${\cal F}$ is a filter on $\omega$. Then ${\cal F}$ has
measure zero or is nonmeasurable. Similarly, it either is meager or
does not have the Baire property.
If ${\cal F}$ is an ultrafilter then ${\cal F}$ is nonmeasurable and
does not have the Baire property. $\vrule width 6pt height 6pt depth 0pt \vspace{0.1in}$
\end{theorem}
In particular if ${\cal F}$ is a filter then $\mu_\star({\cal F})=0$
and if ${\cal F}$ is an ultrafilter then $\mu^\star({\cal F})=1$.
First consider the filters which do not have the Baire property.
\begin{theorem}[Talagrand {[T]}]\label{2.1}
\ \begin{enumerate}
\item The intersection of countably many filters without the
Baire property is
a filter without the Baire property.
\item Assume ${\bf MA}$. Then the
intersection of $< 2^{\aleph_{0}}$ filters
without the Baire property is
a filter without the Baire property. $\vrule width 6pt height 6pt depth 0pt \vspace{0.1in}$
\end{enumerate}
\end{theorem}
For ultrafilters we have a much stronger result.
\begin{theorem}[Plewik {[P]}]\label{plew}
The intersection of $< 2^{\aleph_{0}}$ ultrafilters is a filter which does
not have the
Baire property. $\vrule width 6pt height 6pt depth 0pt \vspace{0.1in}$
\end{theorem}
For Lebesgue measure the situation is more complicated.
\begin{theorem}[Talagrand {[T]}]
Let $\{{\cal F}_n : n \in \omega\}$ be a family of nonmeasurable filters.
Then ${\cal F}=\bigcap_{n \in \omega} {\cal F}_n$ is nonmeasurable.
\end{theorem}
If we consider uncountable families of filters the analog of
\ref{2.1} is no longer true.
\begin{theorem}[Fremlin {[F]}]\label{fr}
Assume Martin's Axiom.
Then there exists a family
$\left\{{\cal F}_{\xi}~ :~ \xi~ <~ 2^{\aleph_{0}}\right\}$ of nonmeasurable filters
such that
$\bigcap_{\xi \in I} {\cal F}_{\xi}$ is a measurable filter
for every uncountable set $I \subset 2^{\aleph_{0}}$.
In particular there exists a family of $\aleph_{1}$ nonmeasurable
filters with measurable intersection.~$\vrule width 6pt height 6pt depth 0pt \vspace{0.1in}$
\end{theorem}
The next theorem shows that the above pathology cannot happen if we assume
stronger measurability properties.
Let $\vec{p} = \<p_{n}:n \in \omega \>$ be a sequence of
reals such that $p_{n}~ \in~ (0,\frac{1}{2}]$ for all $n \in \omega$.
Define $\mu_{\vec{p}}$ to be the product measure on
$2^{\omega}$ such that for all $n$,
$$\mu_{\vec{p}}(\left\{x \in 2^{\omega} : x(n)=1\right\}) =
p_{n}$$ and
$$\mu_{\vec{p}}(\left\{x \in 2^{\omega} : x(n)=0\right\}) = 1
- p_{n}.$$
Notice that if $p_{n} = \frac{1}{2}$ for all $n$ then $\mu_{\vec{p}}$
is the usual measure on $2^{\omega}$.
\begin{theorem}[Bartoszynski {[Ba]}]\label{2.4}
Assume ${\bf MA}$. Let $\mu_{\vec{p}}$ be a
measure such that $\lim_{n \rightarrow \infty} p_{n}=0$ and let
$\left\{{\cal F}_{\xi} :
\xi < \lambda < 2^{\omega}\right\}$ be a family of $\mu_{\vec{p}}$-nonmeasurable
filters.
Then
$$\bigcap_{\xi < \lambda} {\cal F}_{\xi}
\hbox{ is a Lebesgue nonmeasurable filter} .\ \vrule width 6pt height 6pt depth 0pt \vspace{0.1in}$$
\end{theorem}
Finally notice that the additional assumptions in \ref{2.1} and
\ref{fr} are necessary. Namely, we have the following result.
\begin{theorem}
It is consistent with ZFC that there exists a family of filters
${\cal A}$ of size $2^{\aleph_{0}}$ such that
\begin{enumerate}
\item ${\cal A}$ consists of filters which do not have the Baire property
and the intersection of any uncountable subfamily of ${\cal A}$ is
equal to ${\cal F}_0$,
\item ${\cal A}$ consists of non-measurable filters
and the intersection of any uncountable subfamily of ${\cal A}$ is
equal to ${\cal F}_0$.
\end{enumerate}
\end{theorem}
{\sc Proof} \hspace{0.2in}
1) Let ${\bf V}$ be a model of ZFC satisfying CH. Let $\<c^\xi_\eta : \xi <
\omega_1, \eta < \kappa\>$ be a generic sequence of Cohen reals over
${\bf V}$.
Let ${\cal F}_\eta$ be the filter generated by $\left\{c^\xi_\eta : \xi <
\omega_1\right\}$ for $\eta < \kappa$.
One easily checks that this family of filters has the required properties.
2) Use a sequence of random instead of Cohen reals. $\vrule width 6pt height 6pt depth 0pt \vspace{0.1in}$.
\section{Intersection of ultrafilters}
\setcounter{theorem}{0}
In this section we show that the analog of \ref{plew} is not true.
\begin{theorem}\label{main}
It is consistent with ZFC that the intersection of some family of
$<2^{\aleph_{0}}$
ultrafilters has measure zero.
\end{theorem}
{\sc Proof} \hspace{0.2in}
We start with the following observation.
\begin{lemma}
Suppose that $\left\{{\cal F}_\xi : \xi < \kappa\right\}$ is a family of filters
on $\omega$ such that
$\mu_\star(\bigcup_{\xi < \kappa} {\cal F}_\xi)>0 \ .$
Then
$$\mu(\bigcap_{\xi < \kappa} {\cal F}_\xi)=0 \ .$$
Moreover, if ${\cal F}_\xi$'s are ultrafilters, $\bigcap_{\xi <
\kappa} {\cal F}_\xi$ has measure zero iff $\bigcup_{\xi < \kappa} {\cal F}_\xi$
has measure 1.
\end{lemma}
{\sc Proof} \hspace{0.2in}
Suppose that
$\mu_\star(\bigcup_{\xi < \kappa} {\cal F}_\xi)>0$. Since all filters
are assumed to be non-principal we know that
$\mu_\star(\bigcup_{\xi < \kappa} {\cal F}_\xi)=1$. Let $A \subseteq
2^\omega$ be a measure 1 set contained in
$\bigcup_{\xi < \kappa} {\cal F}_\xi$. Define $A^\star = \left\{\omega-X :
X \in A\right\}$. Clearly $\mu(A^\star)=1$.
We claim that
$$A^\star \cap \bigcap_{\xi < \kappa} {\cal F}_\xi = \emptyset \ .$$
Suppose that $X \in A^\star$. Then $\omega-X \in A$ and hence
$\omega-X \in {\cal F}_\eta$ for some $\eta < \kappa$.
It follows that $X \not \in \bigcap_{\xi < \kappa} {\cal F}_\xi$.
Conversely, suppose that $\bigcap_{\xi < \kappa} {\cal F}_\xi$ has measure
zero. Let $A \subseteq 2^\omega$ be a set of measure 1 which is
disjoint with $\bigcap_{\xi < \kappa} {\cal F}_\xi$. If ${\cal F}_\xi$'s are
ultrafilters then it follows that $A^\star \subseteq \bigcup_{\xi <
\kappa} {\cal F}_\xi$.
$\vrule width 6pt height 6pt depth 0pt \vspace{0.1in}$
Let $\left\{I_n : n \in \omega\right\}$ be a partition of $\omega$ into finite
sets such that $|I_n| \geq 2^n$ for $n \in \omega$.
Define
$$A = \left\{X \subseteq \omega : \exists a >1 \
\forall^\infty n \ \frac{|X \cap I_n|}
{|I_n|} > \frac{a}{3}\right\} .$$
It is not hard to see that $\mu(A) = 1$.
Let ${\cal F}$ be an ultrafilter on $\omega$. Define a notion of forcing
${\cal Q}_{\cal F}$ as follows.
$${\cal Q}_{\cal F} = \left\{q \in [A]^{<\omega} : \exists a > 1 \
\left\{n : \frac{|I_n \cap \bigcap q|}{|I_n|} > \frac{a}{3^{|q|}} \right\} \in {\cal F}\right\} .$$
Elements of ${\cal Q}_{\cal F}$ are ordered by inclusion.
\begin{lemma}\label{lem}
${\cal Q}_{\cal F}$ is powerfully ccc.
\end{lemma}
{\sc Proof} \hspace{0.2in}
We have to show that ${\cal Q}^n_{\cal F}$ satisfies countable chain
condition for every natural number $n$.
Let's start with the case when $n=1$.
Suppose that ${\cal A} \subseteq {\cal Q}_{\cal F}$ is an uncountable
subset. Find $k \in \omega$, $a>1$ and an uncountable set
${\cal A}' \subseteq {\cal A}$ such
that $|q|=k$ and
$$ \left\{n : \frac{|I_n \cap \bigcap q|}{|I_n|} \geq \frac{a}{3^k} \right\} \in {\cal F} \ $$
for all $q \in {\cal A}'$.
We show that any sufficiently big subset of ${\cal A}'$
contains two compatible conditions.
We will use the following general observation.
\begin{lemma}
Let $(X, \nu)$ be a measure space with probability measure $\nu$.
Suppose that $1<b<a$ and $\varepsilon>0$. There exists a number $l$
such that if $A_1, \ldots, A_l$ are subsets of $X$ of measure $\geq a
\cdot \varepsilon$ then there are $i \neq j$ such that
$$\nu(A_i \cap A_j) \geq b^2 \cdot \varepsilon^2.$$
\end{lemma}
{\sc Proof} \hspace{0.2in}
Let $A_1, \ldots, A_l$ be sets of measure $a \cdot \varepsilon$.
Consider random variables $X_i$ given by characteristic function of
$A_i$ for $i \leq l$. Note that $X_i^2=X_i$ for $i \leq l$.
Suppose that $\nu(A_i \cap A_j) < b^2 \cdot \varepsilon^2$ for $i
\neq j$. In particular $E(X_i X_j) < b^2 \cdot \varepsilon^2$ for $i
\neq j$. Recall that ${\bf E}(X^2) \geq \lft1({\bf E}(X)\rgt1)^2$ for every
random variable $X$.
We compute
$$\eqalign{
0&\leq{\bf E}\lft2(\lft1((\sum_{i=1}^l X_i)-la\varepsilon\rgt1)^2\rgt2)\cr
&={\bf E}\lft1(\sum_{i=1}^l X_i^2+\sum_{i\neq j}X_iX_j
-2la\varepsilon\sum_{i=1}^l X_i+(la\varepsilon)^2\rgt1)\cr
&\leq la\varepsilon+l(l-1)b^2\varepsilon^2
-2(la\varepsilon)^2+(la\varepsilon)^2\cr
&=l\lft1(a\varepsilon-b^2\varepsilon^2+l(b^2-a^2)\varepsilon^2\rgt1)
.}$$
As $b<a$, the last line here is negative for large l, a contradiction.~$\vrule width 6pt height 6pt depth 0pt \vspace{0.1in}$
Let $\nu_n$ be the uniform measure on $I_n$ i.e. $\nu(A) = |A| \cdot
|I_n|^{-1}$ for $A \subseteq I_n$. Fix $1<b<a$ and let $l$ be a number
from the above lemma chosen for $\varepsilon=3^{-k}$.
Let $q_1, \ldots, q_l \in {\cal A}'$.
Find $Y \in {\cal F}$ such that for all $i \leq l$ and $n \in Y$
$$\frac{|I_n \cap \bigcap q_i|}{|I_n|} \geq
\frac{a}{3^k} \ .$$
By the lemma for every $n \in Y$ there exist $i \neq j$ such that
$$\frac{|I_n \cap \bigcap (q_i \cup q_j)|}{|I_n|} \geq
\frac{b}{9^k} \ .$$
Since ${\cal F}$ is an ultrafilter there exist $i \neq j$ such that
$$\left\{n : \frac{|I_n \cap \bigcap (q_i \cup q_j)|}{|I_n|} \geq
\frac{b}{9^k}\right\} \in{\cal F} \ .$$
It follows that $q_i \cup q_j \in {\cal Q}_{\cal F}$.
Note that in fact we proved that for every $m$ there exists $l$ such
that for every subset $X \subseteq {\cal A}'$ of size $l$ there are
$q_1, \ldots, q_m \in X$ such that $q_i \cup q_j \in {\cal
Q}_{\cal F}$ for $i,j \leq m$.
Suppose that $n>1$. For simplicity assume that $n=2$, the general case
is similar.
Let ${\cal A}$ be an uncountable subset of ${\cal Q}_{\cal F}^2$.
Without loss of generality we can assume there are numbers $k_1$,
$k_2$ and $a$ such that
for every $\<q^1,q^2\> \in {\cal A}$, $|q^1|=k_1$, $|q^2|=k_2$ and
$$ \left\{n : \frac{|I_n \cap \bigcap q^i|}{|I_n|} \geq
\frac{a}{3^{k_i}} \right\} \in {\cal F} \hbox{ for } i=0,1 . $$
To get two elements of ${\cal A}$ which are compatible first apply the
above remark to get a large subset $X \subseteq {\cal A}$ with first
coordinates being pairwise compatible and then apply the case $n=1$ to
the second coordinates of conditions in $X$.~$\vrule width 6pt height 6pt depth 0pt \vspace{0.1in}$
Notice that if $G$ is a ${\cal Q}_{\cal F}$-generic filter over ${\bf V}$ then
the set
$$\left\{\bigcap q : q \in G \right\} \cup {\cal F}_0$$
generates a non-principal filter on $\omega$.
Let ${\cal P}_{\cal F} = \lim_{n \rightarrow \infty} {\cal Q}_{\cal F}^n$ be a finite
support product of countably many copies of ${\cal Q}_{\cal F}$.
By \ref{lem}, ${\cal P}_{\cal F}$ satisfies the countable chain
condition.
\begin{lemma}\label{crucial}
$\mathrel{\|}\joinrel\mathrel{-}_{{\cal P}_{\cal F}} A \cap {\bf V} \hbox{ is the union of countably many
filters} .$
\end{lemma}
{\sc Proof} \hspace{0.2in}
Let $G$ be a ${\cal P}_{\cal F}$-generic filter over ${\bf V}$.
For $n \in \omega$,
let $G_n = \left\{ p(n): {\rm dom}(p)=\left\{n\right\}, p \in G\right\}$.
Since $G_n$ is a ${\cal
Q}_{\cal F}$-generic filter let ${\cal F}_n$ be a filter on $\omega$ generated by
$G_n$ as above.
We show that ${\bf V}[G] \models A \cap {\bf V} \subseteq \bigcup_{n \in \omega}
{\cal F}_n$.
Suppose that $\bar{p} \in {\cal P}_{\cal F}$ and $X \in A$. Find $n \in
\omega$ such that $n \not \in {\rm dom}(\bar{p})$. Let $\bar{q}=\bar{p}
\cup \<n,\left\{X\right\}\>$. It is clear that $\bar{q} \mathrel{\|}\joinrel\mathrel{-} X \in {\cal F}_n$.
$\vrule width 6pt height 6pt depth 0pt \vspace{0.1in}$
Now we can finish the proof of \ref{main}.
Let ${\bf V} \models 2^{\aleph_{0}}>\aleph_1$. Let $\<{\cal P}_\xi, \dot{{\cal
Q}}_\xi : \xi < \omega_1\>$ be a finite support iteration such that
$$\mathrel{\|}\joinrel\mathrel{-}_\xi \dot{{\cal Q}}_\xi \simeq {\cal P}_{{\cal F}_\xi} \hbox{ for
some ultrafilter } {\cal F}_\xi\ .$$
Let ${\cal P}_{\omega_1} = \lim_{\xi < \omega_1} {\cal P}_\xi$.
Let $G$ be a ${\cal P}_{\omega_1}$-generic filter over ${\bf V}$. We will
show that ${\bf V}[G]$ is the model we are looking for.
Let $\left\{{\cal F}^\xi_n : \xi < \omega_1, n \in \omega\right\}$ be the family of
filters added by $G$.
Without loss of generality we can assume that they are ultrafilters.
To finish the proof it is enough to show that
$$A \subseteq \bigcup_{\xi < \omega_1} \bigcup_{n \in \omega} {\cal F}^\xi_n
\ .$$
Suppose that $X \in A$. There exists $\xi < \omega_1$ such that $X \in
{\bf V}[G \cap {\cal P}_\xi]$. By \ref{crucial} there exists $n \in
\omega$ such that $X \in F_n^{\xi +1}$. $\vrule width 6pt height 6pt depth 0pt \vspace{0.1in}$
|
\section{Introduction\ }
Equivariant intersection theory is similar to Borel's equivariant
cohomology. The common basic idea is simple. Let $X$ be an algebraic scheme
over a field $k$ and let $G$ be an algebraic group acting on $X$. Since
invariant cycles are often too few to get a full-fledged intersection theory
(e.g. to have a ring structure in smooth cases) we decide to enlarge this
class to include invariant cycles not only on $X$ but on $X\times V$ where $%
V $ is any linear representation of $G$. If $k=\mathbf{C}$, equivariant
cohomology. can be defined along these lines and this definition agrees with
the usual one given using the classifying space of $G$.
In particular, we get a non trivial equivariant intersection theory $%
A_G^{*}=A_G^{*}(pt)$ on $pt=Speck$ which can be interpreted naturally as an
intersection theory on the classifying stack of the group in the same way as
equivariant cohomology of a point is naturally viewed as cohomology of the
classifying space of the group.
Equivariant intersection theory (in the sense sketched above) was first
defined by Totaro in \cite{To2} for $X=Speck$ and then extended to general $%
X $ by Edidin and Graham in \cite{EG1}. Totaro himself (\cite{To2}) and
Pandharipande (\cite{Pa1}, \cite{Pa2}) computed $A_G^{*}$ in many
interesting cases, for example $G=GL_n,$ $SL_n$ (these two cases are
trivial), $O(n),$ $SO(2n+1)$ and $SO(4)$.
Moreover, Totaro (\cite{To1}, \cite{To2}) was able to define a remarkable
refining of the classical cycle map from the Chow ring to the cohomology
ring. In particular, he proved that for any complex algebraic group $G$, the
equivariant version of the cycle class map $\mathrm{cl}_G:A_G^{*}\rightarrow
H^{*}\left( \mathrm{B}G,\mathbf{Z}\right) $, factors as
\[
A_G^{*}\stackrel{\widetilde{\mathrm{cl}}_G}{\longrightarrow }MU^{*}\left(
\mathrm{B}G\right) \otimes _{MU^{*}}\mathbf{Z}\longrightarrow H^{*}\left(
\mathrm{B}G,\mathbf{Z}\right)
\]
where $MU^{*}\left( \mathrm{B}G\right) $ is the complex cobordism of the
classifying space of $G$,
\[
MU^{*}\left( pt\right) \equiv MU^{*}=\mathbf{Z}\left[ x_1,x_2,x_3,\ldots
\right]
\]
where $\deg x_i=-2i$ and $\mathbf{Z}$ is viewed as a $MU^{*}$-module via the
map sending each $x_i$ to zero. He conjectured that, if $MU^{*}\left(
\mathrm{B}G\right) \otimes _{MU^{*}}\mathbf{Z}$ is concentrated in even
degrees, then $\widetilde{\mathrm{cl}}_G$ is an isomorphism.
The case $G=PGL_n$ is of particular interest. One reason is its connection
with Brauer-Severi varieties, whose Chow groups are quite mysterious (see
\cite{Ka1} and \cite{Ka2} for some results on codimension $2$ cycles). Also,
many parameter spaces of interest are quotients of free actions of $PGL_n$,
so the calculation of $A_{PGL_n}^{*}$ would be a necessary first step to
determine the Chow ring of some of these spaces.
Unfortunately, the ring $A_{PGL_n}^{*}$ for general $n$ seems extremely
difficult to compute. It is a general principle that among all families of
classical groups the series $PGL_n$ is often the hardest to study. Thus, for
example, while the cohomology and the complex cobordism ring of most
classical groups have been determined, very little is known about the
torsion part in the cohomology of the classifying space of $PGL_n$ for $%
n\geq 4$. Of course, given how much harder than cohomology the Chow ring
usually is, this is not encouraging. On the other hand, the cohomology with $%
\mathbf{Z}/3$ coefficients of the classifying space of $PGL_3$, as well as
its Brown-Peterson cohomology (relative to the prime $3$) have been computed
by Kono, Mimura and Shimada (\cite{KMS}) and by Kono and Yagita (\cite{KY}).
The ring $A_{PGL_2}^{*}$ was first computed by Pandharipande (\cite{Pa1})
through the isomorphism $PGL_2\simeq SO\left( 3\right) $. Pandharipande's
method does not seem to extend to $PGL_3$.
In this paper we study $A_{PGL_3}^{*}$. Our approach is completely
different. The idea is that the adjoint representation $\mathrm{sl}_n$ of $%
PGL_n$ can be stratified, using Jordan canonical form, in such a way that
the equivariant Chow ring of each stratum is amenable to study. This
determines completely $A_{PGL_2}^{*}$ (\cite{Ve}) and works fairly well for $%
n=3$ yielding generators of $A_{PGL_3}^{*}$. In principle this method could
give generators for $A_{PGL_n}^{*}$ for any $n$, but the calculations become
extremely involved as $n$ grows. Moreover, as usual, the stratification
method is not very good for finding the relations. In the case $n=3$, using
also a recent general result by Totaro (Th. \ref{gottliebtotaro}), we find
some of the relations in section 5, but unfortunately we are not able to
prove that our relations are sufficient.
We also prove some properties of the cycle map and of Totaro's refined cycle
map. In particular, we are able to prove that $A_{PGL_3}^{*}$, unlike $%
A_{PGL_2}^{*}$, is not generated by Chern classes of representations, a
result conjectured by Totaro in \cite{To2}. We have two proofs of this fact,
one (Th. \ref{result1}), relying on results of \cite{To2}, \cite{KMS}, \cite
{KY}, carries more informations on the cycle and refined cycle maps while
the other (Appendix) is self-contained not depending on cohomological
arguments.
Most of the results in this paper constitute the core of \cite{Ve}.
\smallskip\
Now we state the main results of this work in more detail.
It is already clear ''rationally'', that Chern classes of the adjoint
representation alone do not generate $A_{PGL_3}^{*}$. So, if $E$ is the
standard representation of $GL_3$, we also consider $Sym^3E$, the $PGL_3$%
-representation defined by
\[
\left[ g\right] \cdot \left( v_1\cdot v_2\cdot v_3\right) \doteq \det
g^{-1}\left( gv_1\cdot gv_2\cdot gv_3\right) \text{.}
\]
We prove the following (Theorem \ref{generators}):
\begin{theorem}
\label{duos}There exist elements $\rho $ and $\chi $ with\emph{\ }$\deg \rho
=4$\emph{\ }and $\deg \chi =6$, such that $A_{PGL_3}^{*}$ is generated by
\[
\left\{ \lambda \doteq 2c_2\left( \mathrm{sl}_3\right) -c_2\left(
Sym^3E\right) ,c_3\left( Sym^3E\right) ,\rho ,\chi ,c_6\left( \mathrm{sl}%
_3\right) ,c_8\left( \mathrm{sl}_3\right) \right\} \text{.}
\]
\end{theorem}
The question of determining all the relations between this generators is
hard. In this direction, we can prove the following (Th. \ref{relations}):
\begin{proposition}
\label{tres}The generators above satisfy
\[
3\rho =3\chi =3c_8\left( \mathrm{sl}_3\right) =0
\]
\[
\rho ^2=c_8\left( \mathrm{sl}_3\right)
\]
\[
3\left( 27c_6\left( \mathrm{sl}_3\right) -c_3\left( Sym^3E\right)
^2-4\lambda ^3\right) =0\text{.}
\]
\end{proposition}
Moreover, if $R^{*}$ denotes the ring
\[
\frac{\mathbf{Z}\left[ \lambda ,c_3\left( Sym^3E\right) ,\rho ,\chi
,c_6\left( \mathrm{sl}_3\right) ,c_8\left( \mathrm{sl}_3\right) \right] }{%
\frak{R}}
\]
where $\frak{R}$ is the ideal generated by the relations in Prop. \ref{tres}
and\emph{\ }$\deg \rho =4$\emph{, }$\deg \chi =6$, we have (Theorem \ref
{riassunto})
\begin{theorem}
The composition
\[
R^{*}\longrightarrow A_{PGL_3}^{*}\stackrel{\widetilde{\mathrm{cl}}}{%
\longrightarrow }MU^{*}\left( \mathrm{B}PGL_3\right) \otimes _{MU^{*}}%
\mathbf{Z}
\]
is surjective and its kernel is $3$-torsion.
\end{theorem}
Note that this also proves that the $R^{*}\left[ \frac 13\right] \simeq
A_{PGL_3}^{*}\left[ \frac 13\right] $. We also prove that while $\rho $ is
nonzero in cohomology, $\chi $ is zero in cohomology. Thus, by Remark \ref
{totaro}, we also have $\widetilde{\mathrm{cl}}\left( \chi \right) =0$. Note
that if one was able to prove that $\chi \neq 0$ then Totaro's conjecture
would be false. However, despite many efforts, we still do not know whether $%
\chi $ is zero or not.
By a result of Kono and Yagita (\cite{KY}), Totaro's conjecture predicts
that $\widetilde{\mathrm{cl}}$ is actually an isomorphism. We are able to
show that the generator $\rho $ of Theorem \ref{duos} is not in the Chern
subring\footnote{%
I.e. in the subring generated by Chern classes of representations.} of $%
A_{PGL_3}^{*}$, thus proving the following consequence of Totaro's
conjecture (Theorem \ref{result1}):
\begin{theorem}
$A_{PGL_3}^{*}$ is not generated by Chern classes.
\end{theorem}
This same result is proved in the Appendix without using cohomology
computations.
\bigskip\
\textbf{Conventions and notations. }The word ''scheme'' will most of the
time mean ''algebraic scheme over a field $k$''. In section 1, where we try
to give some of the results in greater generality, we will allow a different
base scheme $S$ and the finiteness conditions needed will be properly
specified.
We freely use the functorial point of view for schemes and group schemes
(e.g. \cite{DG}) to be able to express maps, actions etc. as sending
''elements to elements''.
If $s$ is a section of a vector bundle, we denote by $Z\left( s\right) $ its
zero scheme.
Algebraic groups over a field $k$ will always be linear. If $G$ is an
algebraic group over a field $k$, $T_G$ (or simply $T$ if no confusion is
possible) denotes a maximal torus of $G$ and $\widehat{T_G}$ its character
group.
If $\varphi :G\rightarrow H$ is a morphism of algebraic groups over a field $%
k$ and $V$ is a representation of $H$, we denote by $V_{\left( \varphi
\right) }$ or $V_{\left( G\right) }$ the obvious associated $G$%
-representation.
If $E$ denotes the standard $GL_3$-representation, $Sym^3E$ becomes a $PGL_3$%
-representation via
\[
\left[ g\right] \cdot \left( v_1\cdot v_2\cdot v_3\right) \doteq \det
g^{-1}\left( gv_1\cdot gv_2\cdot gv_3\right) \text{.}
\]
\smallskip\
\textbf{Acknowledgments.} I wish to thank my thesis advisor Angelo Vistoli
for his friendship, patience and constant attention to this work.
I also wish to thank Burt Totaro for generous advice and for many
illuminating discussions we had in Cambridge. Among other things, he also
explained to me the argument in Remark \ref{totaro} and allowed me to
include his still unpublished result Theorem \ref{gottliebtotaro}. I hope
this paper shows all its debts to his deep and original work.
\section{ Basic notations and results}
In this section we mainly fix notations and collect some miscellaneous
results on equivariant Chow groups we will need in the sequel; most of them
(with one possible exception) are elementary or well known but we simply
could not find proper references in the literature. For intersection theory
the standard refernce is \cite{Fu} while for equivariant intersection theory
we refer to \cite{EG1} and \cite{To2}.
\subsection{Equivariant intersection theory and Totaro's refined cycle map\ }
If $G$ be an algebraic group over a field $k$ and $X$ a smooth\footnote{%
We restrict our attention to smooth schemes for simplicity.} $G$-scheme.
Edidin and Graham (\cite{EG1}), following an idea of Totaro (\cite{To2}),
defined a $G$-equivariant version, $A_G^{*}\left( X\right) $, of the Chow
ring $A^{*}\left( X\right) $. \label{BG}We will simply write $A_G^{*}$ for $%
A_G^{*}\left( Speck\right) $. As a rule, if we do not mention explicitly the
base field $k$, we are assuming $k=\mathbf{C}$.
We say that a pair $\left( U,V\right) $, consisting of a $k$-representation $%
V$ of $G$ and an open subset $U$ of $V$ on which $G$ acts freely, is a good
pair (or simply a pair) relative to $G$ if the codimension of $V\backslash U$
has sufficiently high codimension (see \cite{EG1}, 2.2,
Definition-Proposition 1).
All the basic properties and constructions (Chern classes, localization
sequence, proper pushforwards, Gysin maps, vector and projective bundle
theorems, projection formula, self intersection formula, cycle class map,
operational Chow groups etc.) of ordinary intersection theory (\cite{Fu})
have their equivariant counterparts. Moreover, there are additional
constructions one can do in the equivariant setting which simply do not
exist in the ordinary case, for example those related to morphisms of
algebraic groups. If
\[
\varphi :G\longrightarrow G^{\prime }
\]
is a morphism of algebraic groups and $X$ a $G^{\prime }$-scheme (which we
suppose smooth just in order to state each result for Chow rings), then $X$
is a $G$-scheme via $\varphi $ and if $\left( U,V\right) $ (respectively, $%
\left( U^{\prime },V^{\prime }\right) $) is a good pair relative to $G\,$%
(resp., relative to $G^{\prime }$), we let $G$ act on $V\times V^{\prime }$
as
\[
g\cdot \left( v,v^{\prime }\right) =\left( g\cdot v,\pi \left( g\right)
\cdot v^{\prime }\right) ,\quad g\in G,\text{ }v\in V,\text{ }v^{\prime }\in
V^{\prime }
\]
and the projection
\[
X\times U\times U^{\prime }\longrightarrow X\times U^{\prime }
\]
induces a flat map
\[
\left( X\times U\times U^{\prime }\right) /G\longrightarrow \left( X\times
U^{\prime }\right) /G^{\prime }.
\]
Its pullback induces a restriction ring morphism
\[
A_{G^{\prime }}^{*}\left( X\right) \longrightarrow A_G^{*}\left( X\right)
\]
denoted by $\varphi _X^{*}$ (or by $\mathrm{res}_{G^{\prime },X}^G$ if $%
\varphi $ is injective). Note that the same construction made in the
topological case, define the functoriality in $G$ of the equivariant
cohomology ring $H_G^{*}\left( X;\mathbf{Z}\right) .$
Another construction which appears only in the equivariant setting is the
following transfer construction for Chow groups; we will frequently use it.
Let
\[
1\rightarrow H\stackrel{\phi }{\longrightarrow }G\longrightarrow
F\rightarrow 1
\]
be an exact sequence of algebraic groups over a field $k$,$\,$with $F$
finite. If $X$ is an algebraic smooth $G$-scheme then $p_1:X\times
F\longrightarrow X$ is proper $G$-equivariant and there is an equivariant
push-forward
\[
p_{1*}:A_{*}^G\left( X\times F\right) \rightarrow A_{*}^G\left( X\right) .
\]
If $\left( U,V\right) $ is a good pair for $G$, we have:
\begin{equation}
\frac{\left( X\times F\right) \times U_{}}G\simeq \left( \frac{\left(
X\times F\right) \times U}H\right) \diagup F\simeq \nonumber
\end{equation}
\[
\simeq \left( \frac{X\times U}H\times F\right) \diagup F\simeq \frac{X\times
U}H;
\]
hence $A_G^{*}\left( X\times F\right) \simeq A_H^{*}\left( X\right) $ and $%
p_{1*}$ induces a \emph{transfer} morphism of graded groups
\begin{equation}
\mathrm{tsf}_{H,X}^G:A_H^{*}\left( X\right) \rightarrow A_G^{*}\left(
X\right) . \label{tsf}
\end{equation}
which is natural in $X$ with respect to pullbacks.
Observe that the pullback $A_G^{*}\left( X\right) \rightarrow A_H^{*}\left(
X\right) $ has actually values in the $F$-invariant subring of $%
A_H^{*}\left( X\right) $ (\cite{To2})
\[
\mathrm{res}_{G,X}^H:A_G^{*}\left( X\right) \longrightarrow \left(
A_H^{*}\left( X\right) \right) ^F.
\]
In exactly the same way as for group cohomology (e.g. \cite{B}, Prop. 9.5),
we have
\[
\mathrm{tsf}_{H,X}^G\circ \mathrm{res}_{G,X}^H=\left( \#F\right) \cdot
\]
(by projection formula) and, since $H$ is normal in $G$,
\begin{equation}
\left( \mathrm{res}_{G,X}^H\circ \mathrm{tsf}_{H,X}^G\right) _{\mid \left(
A_H^{*}\left( X\right) \right) ^F}=\left( \#F\right) \cdot \label{media1}
\end{equation}
If we do not restrict to $\left( A_H^{*}\left( X\right) \right) ^F$, we get
\begin{equation}
\mathrm{res}_{G,X}^H\circ \mathrm{tsf}_{H,X}^G\left( \xi \right) =\sum_{f\in
F}f_{*}\xi \label{media2}
\end{equation}
for any $\xi $ in $A_H^{*}\left( X\right) $.
\begin{remark}
For a general action of $G$ on $X$, the quotient $\left[ X/G\right] $ exists
only as an Artin stack\footnote{%
Not necessarily separated.} (\cite{LMB}). Edidin and Graham (\cite{EG1},
5.3, Prop. 16, 17) showed that if $\mathcal{F}$ is a quotient Artin stack $%
\mathcal{F}\simeq \left[ X/G\right] $, then the corresponding equivariant
Chow groups do not depend on the presentation chosen for the quotient,
enabling one to \emph{define }$A_G^{*}\left( X\right) $ to be the (integral)
Chow group of the stack $\mathcal{F}$. If moreover $\mathcal{F}$ is smooth,
there is a ring structure on this Chow group and this applies to the
classifying stack $\mathcal{B}G$ of any algebraic group $G$ (\cite{LMB}),
viewed as the quotient $\left[ \mathrm{pt}/G\right] $,
\[
A_G^{*}=A^{*}\left( \mathcal{B}G\right) \text{. }
\]
\end{remark}
\begin{theorem}
\label{gottliebtotaro}(Gottlieb; Totaro) Let $G$ be an algebraic group over $%
\mathbf{C}$, $T$ a maximal torus of $G$ and $N_G\left( T\right) $ its
normalizer in $G$. The restriction maps
\begin{equation}
A_G^{*}\longrightarrow A_{N_G\left( T\right) }^{*} \label{totaro}
\end{equation}
\begin{equation}
H^{*}\left( \mathrm{B}G,\mathbf{Z}\right) \longrightarrow H^{*}\left(
\mathrm{B}N_G\left( T\right) ,\mathbf{Z}\right) \label{gottlieb}
\end{equation}
are injective.
\end{theorem}
\TeXButton{Proof}{\proof} (\ref{gottlieb}) is proved in \cite{Gtl}. We
sketch the proof of (\ref{totaro}) from \cite{To3}. If $f:Y\longrightarrow B$
is a smooth proper morphism of relative dimension $r$ between smooth,
separated schemes of finite type over $k$, let us consider the following
''modified'' pushforward
\[
f_{\sharp }\left( \alpha \right) \doteq f_{*}\circ \left( c_r\left( \mathcal{%
T}_f\right) \cdot \alpha \right) \in A^j\left( S\right)
\]
for any $\alpha \in A^j\left( B\right) $, where $\mathcal{T}_f$ denotes the
relative tangent bundle; by projection formula, we have
\begin{equation}
f_{\sharp }\circ f^{*}=\chi \left( F\right) \label{formula}
\end{equation}
where $\chi \left( F\right) $ denotes the Euler characteristic of ''the
fiber'' of $f$ (equal to the degree of the top Chern class of its tangent
bundle). Now, let $g:X\rightarrow B$ be a smooth morphism between smooth
schemes over a field $k$ which admits a smooth relative compactification
\[
\begin{tabular}{lll}
$X$ & $\hookrightarrow $ & $\overline{X}$ \\
& $\searrow $ & $\downarrow $ \\
& & $B$%
\end{tabular}
\]
having divisors with normal crossing $\left\{ D_i\right\} _{i=1,\ldots ,n}$
as complement (smooth over $B$). If $D_{ij}\doteq D_i\cap D_j$, $%
D_{ijk}\doteq D_i\cap D_j\cap D_k$, etc., the previous construction yields
modified pushforwards
\[
\text{ }f_{\sharp }:A^{*}\left( \overline{X}\right) \rightarrow A^{*}\left(
B\right) ,\text{ }f_{\sharp }^{\left( 1\right) }:\oplus _iA^{*}\left(
D_i\right) \rightarrow A^{*}\left( B\right) ,\text{ }f_{\sharp }^{\left(
2\right) }:\oplus _{i<j}A^{*}\left( D_{ij}\right) \rightarrow A^{*}\left(
B\right) ,\ldots
\]
satisfying (\ref{formula}). If $x\in A^{*}\left( X\right) $, lift it to some
$\overline{x}\in A^{*}\left( \overline{X}\right) $ and set
\[
g_{\sharp }\left( x\right) \doteq f_{\sharp }\left( \overline{x}\right)
-\sum_if_{\sharp }^{\left( 1\right) }\left( \overline{x}_{\mid D_i}\right)
+\sum_{i<j}f_{\sharp }^{\left( 2\right) }\left( \overline{x}_{\mid
D_{ij}}\right) -\cdots
\]
(alternating sum) which is an element in $A^{*}\left( B\right) $. This can
be shown to be independent on the choice of the lifting and (\ref{formula})
holds for $g$ by well-known properties of the Euler characteristic.
To prove (\ref{totaro}), apply this construction to any approximation of the
$G/N_G\left( T\right) $-torsor
\[
\mathrm{B}N_G\left( T\right) \longrightarrow \mathrm{B}G
\]
recalling that $\chi \left( G/N_G\left( T\right) \right) =1$. Note that this
proof works over any algebraically closed field $k$. \TeXButton{End Proof}
{\endproof}
In \cite{To2}, Totaro proved the remarkable fact that, if $G$ is a complex
algebraic group, the cycle map
\[
\mathrm{cl}_{\mathrm{B}G}:A_G^{*}\longrightarrow H^{*}\left( \mathrm{B}G,%
\mathbf{Z}\right)
\]
factors as
\begin{equation}
A_{\mathrm{B}G}^{*}\stackrel{\widetilde{\mathrm{cl}}_{\mathrm{B}G}}{%
\longrightarrow }MU^{*}\left( \mathrm{B}G\right) \otimes _{MU^{*}}\mathbf{Z}%
\stackrel{\underline{\mathrm{cl}}_{\mathrm{B}G}}{\longrightarrow }%
H^{*}\left( \mathrm{B}G,\mathbf{Z}\right) . \label{factor}
\end{equation}
where $MU^{*}\left( \mathrm{B}G\right) $ is the complex cobordism ring of $%
\mathrm{B}G$ (\cite{St}) and $\underline{\mathrm{cl}}_{\mathrm{B}G}$ is the
natural morphism (since
\[
MU^{*}\equiv MU^{*}\left( \mathrm{pt}\right) =\mathbf{Z}\left[
x_1,x_2,\ldots ,x_n,\ldots \right]
\]
with $\deg x_i=-2i$, here $\mathbf{Z}$ is viewed as an $MU^{*}$-module via
the map sending each generator $x_i$ to zero). We call $\widetilde{\mathrm{cl%
}}_{\mathrm{B}G}$ \emph{Totaro's refined cycle map} for $G$. The kernel and
cokernel of $\underline{\mathrm{cl}}_{\mathrm{B}G}$, $\widetilde{\mathrm{cl}}%
_{\mathrm{B}G}$ and $\mathrm{cl}_{\mathrm{B}G}$ are torsion.
In \cite{To2}, Totaro conjectures that if $G$ is a complex algebraic group
such that $MU^{*}\left( \mathrm{B}G\right) $, localized at some prime $p$,
is concentrated in even degrees, then the $p$-localization of $\widetilde{%
\mathrm{cl}}_{\mathrm{B}G}$ should be an isomorphism. As a consequence of
this conjecture, $A_{PGL_3}^{*}$ should not be generated by Chern classes
since, by \cite{KY}, $MU^{*}\left( \mathrm{B}PGL_3\right) $ is concentrated
in even degrees but not generated by Chern classes. This consequence of
Totaro's conjecture will be proved in section 4 (see also the Appendix for a
different proof).
\begin{remark}
Voevodsky (\cite{V1}, \cite{V2}) defined an \emph{algebraic cobordism}%
\textrm{\ }for an algebraic scheme over an \emph{arbitrary} field $k$, so it
would be interesting to investigate if there exists a generalization of
Totaro's refined cycle map with values in (a quotient of) algebraic
cobordism, for any algebraic group $G$ over $k$. As M. Levine suggested to
me, one may also ask more generally if Totaro's refined cycle map extends to
a map from the entire Voevodsky's motivic cohomology to algebraic cobordism.
\end{remark}
\subsection{Miscellaneous results}
\begin{proposition}
\label{rational} Let $k$ be algebraically closed. The pullback $%
A_{PGL_{n,k}}^{*}\otimes \mathbf{Q}\rightarrow A_{SL_{n,k}}^{*}\otimes
\mathbf{Q}$ is an isomorphism.
\end{proposition}
\TeXButton{Proof}{\proof} By \cite{EG2}, Th. 1 (c),
\[
A_G^{*}\otimes \mathbf{Q}\simeq \mathrm{Sym}_{\mathbf{Z}}\left( \widehat{T}%
\right) ^W\otimes \mathbf{Q=}\left( \mathbf{A}_T^{*}\right) ^W\otimes
\mathbf{Q}
\]
for any connected reductive algebraic group $G$ with maximal torus $T$ and
Weyl group $W$ and $\mathrm{Sym}_{\mathbf{Z}}\left( \widehat{T}\right)
^W\otimes \mathbf{Q}$ is the same for a group $G$ and a quotient of $G$ by a
finite central subgroup. \TeXButton{End Proof}{\endproof}
\begin{remark}
\label{equivalence}Let $S$ be a locally noetherian base scheme. Since $%
\mathrm{Aut}\left( \mathbf{P}_S^n\right) \simeq PGL_{n+1,S}$ as
group-functors, for any $S$ (\cite{DG} or \cite{GIT}, p. 20-21), the
category of Brauer-Severi schemes (\cite{Mi}, p. 134) of relative dimension $%
n$ over $X$ for the \'etale (or \textrm{fppf}) topology is equivalent to
that of $PGL_{n+1}$-torsors over $X$ for the same topology and this
equivalence actually extends to a $1$-isomorphism of $\mathcal{BS}_{n,S}$
with the classifying stack $\mathcal{B}\left( PGL_{n+1,S}\right) $, where $%
\mathcal{BS}_{n,S}$ denotes the stack over $S$ whose fibre category over $%
X/S $ is the category of Brauer-Severi schemes of relative dimension $n$
over $X$. Under this $1$-isomorphism trivial\footnote{%
I.e. of the form $P\left( E\right) \rightarrow X$ for some vector bundle $E$
over $X$.} Brauer-Severi schemes correspond to $PGL_{n+1}$-torsors induced
by $GL_{n+1}$-torsors.
\end{remark}
\begin{proposition}
Let $k$ be algebraically closed. Then \label{torsion}$\ker
(A_{PGL_{n,k}}^{*}\rightarrow A_{SL_{n,k}}^{*})$ is $n$-torsion.
\end{proposition}
\TeXButton{Proof}{\proof} By Prop. \ref{rational}, our kernel is torsion and
so it is enough to prove that $\ker (p^{*}:A_{PGL_{n,k}}^{*}\rightarrow
A_{GL_{n,k}}^{*})$ is annihilated by $n$, $A_{GL_{n,k}}^{*}$ being torsion
free.
By \cite{To2}, Th. 1.3 or \cite{EG2} Th. 1, for any reductive algebraic
group $G$, $A_G^{*}$ can be identified with the ring $\mathcal{C}_G^{*}$ of
characteristic classes for (\'etale) $G$-torsors over smooth, separated
schemes of finite type over $k$. Via this identification $p^{*}$ translates
to
\begin{eqnarray*}
p^{*} &:&\mathcal{C}_{PGL_{n.k}}^{*}\longrightarrow \mathcal{C}%
_{GL_{n,k}}^{*} \\
F &\longmapsto &p^{*}(F):\left(
\begin{array}{c}
E \\
\downarrow \\
X
\end{array}
\right) \mapsto F\left(
\begin{tabular}{l}
$P_E$ \\
$\downarrow $ \\
$X$%
\end{tabular}
\right)
\end{eqnarray*}
where $P_E\rightarrow X$ is the $PGL_{n,k}$-torsor associated to $\mathbf{P}(%
\widetilde{E})\rightarrow X$, $\widetilde{E}\rightarrow X$ being the vector
bundle associated to the $GL_{n,k}$-torsor $E\rightarrow X$ and slightly
abusing notation in the argument of $F$
\[
p^{*}F=0\text{ }\Longleftrightarrow \text{ }F(\mathbf{P}(\widetilde{E}%
)\rightarrow X)=0,\text{ }\forall E\rightarrow X\text{ vector bundle of }rk%
\text{ }n.
\]
Now we use the $1$-isomorphism of stacks $\mathcal{B}\left( PGL_{n,k}\right)
\simeq \mathcal{BS}_{n-1,k}$ (Remark \ref{equivalence}). If $f:P\rightarrow
X $ is a $PGL_{n,k}$-torsor and $\overline{f}:\overline{P}\rightarrow X$ the
associated Brauer-Severi scheme, the base change of $f$ via $\overline{f}$
is a $PGL_{n,k}$-torsor induced by a $GL_{n,k}$-torsor. Since $\chi \left(
P_k^{n-1}\right) =n$, formula (\ref{formula}) in the proof of Theorem \ref
{gottliebtotaro} yields
\[
nF\left(
\begin{tabular}{l}
$P$ \\
$\downarrow $ \\
$X$%
\end{tabular}
\right) =\overline{f}_{\sharp }\overline{f}^{*}F\left(
\begin{tabular}{l}
$P$ \\
$\downarrow $ \\
$X$%
\end{tabular}
\right) =\overline{f}_{\sharp }F(\overline{f}^{*}\left(
\begin{tabular}{l}
$P$ \\
$\downarrow $ \\
$X$%
\end{tabular}
\right) )=0
\]
(by projection formula) if $p^{*}F=0$. \TeXButton{End Proof}{\endproof}
\begin{corollary}
\label{cortorsion}$A_{PGL_{n,k}}^{*}$ has only $n$-torsion.
\end{corollary}
We conclude this section collecting some elementary results on equivariant
Chow groups we will use in the sequel.
\begin{proposition}
$A_{\mathbf{\mu }_{n,k}}^{*}\simeq \mathbf{Z}\left[ t\right] \diagup \left(
nt\right) $.
\end{proposition}
\TeXButton{Proof}{\proof} From Kummer exact sequence
\[
1\rightarrow \mathbf{\mu }_{n,k}\longrightarrow \mathbf{G}_{m,k}\stackrel{(%
\text{ })^n}{\longrightarrow }\mathbf{G}_{m,k}\rightarrow 1,
\]
for any $N>0$ we get a $\mathbf{G}_{m,k}$-torsor
\[
\frac{\mathbf{A}_k^{N+1}\backslash \left\{ 0\right\} }{\mathbf{\mu }_{n,k}}%
\longrightarrow \frac{\mathbf{A}_k^{N+1}\backslash \left\{ 0\right\} }{%
\mathbf{G}_{m,k}}=\mathbf{P}_k^N
\]
whose associated line bundle is just $\mathcal{O}_{\mathbf{P}_k^N}(-n)$. By
\cite{Gr1}, Remark p. 4-35, we get
\[
A^{*}\left( \frac{\mathbf{A}_k^{N+1}\backslash \left\{ 0\right\} }{\mathbf{%
\mu }_{n,k}}\right) \simeq \frac{A^{*}\left( \mathbf{P}_k^N\right) }{\left(
c_1\left( \mathcal{O}_{\mathbf{P}_k^N}(-n)\right) \right) }
\]
which implies the assert for $N\gg 0$. \TeXButton{End Proof}{\endproof}
\begin{proposition}
\label{unipotent}If $G$ is a unipotent algebraic group over a field $k$ of
characteristic zero, then $A_G^{*}\simeq \mathbf{Z}$.
\end{proposition}
\TeXButton{Proof}{\proof}Since $G$ is unipotent it has a central composition
series
\[
G=G_0 > G_1 > G_2 > \cdots
> G_n=1
\]
such that $G_i/G_{i+1}\simeq \mathbf{G}_{a,k}$ (\cite{DG}, IV, \S\ 2, 2.5
(vii)). We proceed by induction on the length $n$ of the composition series.
If $n=1$, $G\simeq \mathbf{G}_{a,k}$; if $U$ is a $G$-free open subset of a $%
G$-representation such that $\pi :U\rightarrow U/G$ is a (\textrm{fppf} or
\'etale) $G$-torsor then $\pi $ is a Zariski $G$-torsor ($\mathbf{G}_{a,k}$
being special, \cite{Se}) and in particular a Zariski affine bundle with
fiber $\mathbf{A}_k^1$ so that $\pi ^{*}$ is an isomorphism (\cite{Gr1}, p.
4-35).
Suppose the assert true for any unipotent group whose central composition
series has length $\leq n$. If $G$ is unipotent with a central decomposition
series
\[
G=G_0 > G_1 > G_2 > \cdots
> G_{n+1}=1
\]
then $G_1$ is unipotent (\cite{DG}, IV, \S\ 2, 2.3) and we have a short
exact sequence
\[
1\rightarrow G_1\longrightarrow G\longrightarrow G/G_1\simeq \mathbf{G}%
_{a,k}\rightarrow 1.
\]
Therefore, if $U$ is a $G$-free open subset of a $G$-representation which
has a $G$-torsor quotient $U\rightarrow U/G$,
\[
U/G_1\rightarrow U/G
\]
is a $G/G_1\simeq \mathbf{G}_{a,k}^{}$-torsor. As in case $n=1$, the
pullback is an isomorphism $A_G^{*}\simeq A_{G_1}^{*}$ and we conclude since
$G_1$ has a central decomposition series of length $n$. \TeXButton{End Proof}
{\endproof}
\begin{proposition}
\label{corunip}Let
\begin{equation}
1\rightarrow H\longrightarrow G\stackrel{\rho }{\stackunder{\sigma }{%
\rightleftarrows }}\mathbf{G}_m\rightarrow 1 \label{extunip}
\end{equation}
be a split exact sequence of algebraic groups over a field $k$ of
characteristic zero, with $H$ unipotent. Then the pullback
\[
\rho ^{*}:A_{\mathbf{G}_m}^{*}\longrightarrow A_G^{*}
\]
is an isomorphism.
\end{proposition}
\TeXButton{Proof}{\proof} Let $U$ be a $G$-free open subset of a $G$%
-representation with complement of sufficiently high codimension and with a $%
G$-torsor quotient $U\rightarrow U/G$. Then
\[
U/H\rightarrow U/G
\]
is a $\mathbf{G}_m$-torsor which corresponds to some line bundle $L$ over $%
U/G$ (Th. \ref{equivalence}) and by \cite{Gr1}, Remark p. 4-35,
\[
A^{*}\left( U/H\right) \simeq \frac{A^{*}\left( U/G\right) }{c_1(L)}.
\]
Since $A_H^{*}\simeq \mathbf{Z}$, by Proposition (\ref{unipotent}), $A_G^{*}$
is then generated by $c_1(L)$. But the pullback $\mathbf{Z}[u]\simeq A_{%
\mathbf{G}_{m,k}}^{*}\rightarrow A_G^{*}$ sends $u$ to $c_1(L)$, therefore $%
\rho ^{*}$ is surjective. Injectivity follows from the hypothesis that (\ref
{extunip}) is split. \TeXButton{End Proof}{\endproof}
\begin{proposition}
\label{productbygm}If $G$ is an algebraic group over $k$, then $A_{G\times
\mathbf{G}_{m,k}}^{*}\simeq A_G^{*}\otimes A_{\mathbf{G}_{m,k}}^{*}$.
\end{proposition}
\TeXButton{Proof}{\proof} Straightforward using $\left( \mathbf{A}%
_k^{N+1}\backslash \left\{ 0\right\} ,\mathbf{A}_k^{N+1}\right) $ as a good
pair for $G_{m,k}$, $N\gg 0$, and \cite{Fu}, Example 8.3.7.
\TeXButton{End Proof}{\endproof}
\begin{proposition}
\label{homogeneous}Let $G$ be an algebraic group over $k$. If $H$ is a
closed algebraic subgroup of $G$, then there is a canonical isomorphism $%
A_G^{*}\left( G/H\right) \simeq A_H^{*}$.
\end{proposition}
\TeXButton{Proof}{\proof} Straightforward. \TeXButton{End Proof}{\endproof}
\begin{proposition}
\label{utrivial}Let $G$ be an algebraic group over a field $k$ and $X$ a
smooth $G$-scheme. If $U\subset \mathbf{A}_k^n$ is an open subscheme with
the trivial $G$-action, the pull-back $pr_2^{*}:A_G^{*}\left( X\right)
\simeq A_G^{*}\left( U\times X\right) $ is an isomorphism .
\end{proposition}
\TeXButton{Proof}{\proof} Since $G$ acts trivially on $U$, we can reduce to
the case of trivial $G$. By \cite{Fu}, Prop. 1.9, the pull back via $\mathbf{%
A}_k^n\times X\rightarrow X$ is surjective and so is $pr_2^{*}$ by the
localization exact sequence (\cite{Fu} Prop. 1.8).
If $k$ is infinite then $pr_2$ has always a section so that $pr_2^{*}$ is
injective. If $k$ is finite, let $p\in U$ be a closed point with $r\doteq $ $%
\left[ k(p):k\right] $. From the commutative diagram
\[
\begin{tabular}{lll}
& $\quad U\times X$ & \\
& $\nearrow \quad \downarrow ^{pr_2}$ & \\
$p\times X$ & $\stackunder{\phi }{\longrightarrow }X$ &
\end{tabular}
\]
and projection formula we get that $\ker \left( pr_2^{*}\right) $ is $r$%
-torsion. Now observe that we can always find two closed points $p$ and $%
p^{\prime }$ in $U$ with residue fields of relatively prime degrees $r$ and $%
r^{\prime }$ over $k$, so that $\ker \left( pr_2^{*}\right) $ is indeed
zero. \TeXButton{End Proof}{\endproof}
\begin{proposition}
\label{change}Let $G,H$ be algebraic groups having commuting actions on a
smooth scheme $X$ and suppose $G$ acts freely. Then there is a canonical
isomorphism $A_H^{*}\left( X/G\right) \simeq A_{G\times H}^{*}\left(
X\right) $.
\end{proposition}
\TeXButton{Proof}{\proof} If $\left( U,V\right) $ is a good pair for $H$,
with $\mathrm{codim}\left( V\backslash U\right) >i$, we have
\[
A_H^i\left( X\diagup G\right) \simeq A^i\left( \left( U\times \frac
XG\right) \diagup H\right) \simeq
\]
\[
\simeq A^i\left( \left( U\times X\right) \diagup G\times H\right) \simeq
A_{G\times H}^i\left( X\times U\right) ,\text{ }
\]
by \cite{EG1}, Prop. 8. By the localization sequence,
\[
A_{G\times H}^i\left( X\times U\right) \simeq A_{G\times H}^i\left( X\times
V\right)
\]
for $i-$\textrm{codim}$(V\backslash U)<0$ and we conlude since for any $%
G\times H$-representation $E$, we have a pullback ring isomorphism $%
A_{G\times H}^{*}\left( X\right) \simeq A_{G\times H}^{*}\left( X\times
E\right) $.\TeXButton{End Proof}{\endproof}
\section{ Generators for $A_{PGL_3}^{*}$}
From now on, our base field will be $\mathbf{C}$.
By Prop. \ref{rational}, we have
\[
A_{PGL_3}^{*}\otimes \mathbf{Q}\simeq A_{SL_3}^{*}\otimes \mathbf{Q=Q}\left[
c_2\left( E\right) ,c_3\left( E\right) \right]
\]
($E=$ standard representation of $SL_3$) and an easy computation shows that $%
c_3\left( E\right) $ is not in the image of the subring of $%
A_{PGL_3}^{*}\otimes \mathbf{Q}$ generated by the Chern classes of $\mathrm{%
sl}_3$. Therefore the Chern classes of the adjoint representation will
certainly not suffice to generate $A_{PGL_3}^{*}$.
In this section we find generators of $A_{PGL_3}^{*}$ (Prop. \ref{prelgen})
by stratifying the adjoint representation $\mathrm{sl}_3$ using Jordan
canonical forms.
Let $G$ be a complex algebraic group. For our purposes a finite $G$%
-stratification of a $G$-scheme $X$ will be a collection $\left\{
X_i\right\} _{i=1,\ldots ,n}$ of disjoint smooth $G$-invariant subschemes,
whose union is $X$ and such that for each $i$ the natural immersion
\[
j_i:X_i\hookrightarrow X\diagdown \bigcup_{k>i}X_k\doteq X^i
\]
is closed. In particular, $X_n$ is a closed subscheme of $X$, each $X_i$ is
topologically a locally closed subspace of $X$ and all the maps
\[
X_1=X^1\hookrightarrow X^2\hookrightarrow X^3\hookrightarrow \cdots
\hookrightarrow X^{n-1}\hookrightarrow X^n\hookrightarrow X
\]
are open immersions. Any stratification $\left\{ X_i\right\} _{i=1,\ldots
,n} $ gives then rise to the following exact sequences (of graded abelian
groups, $\deg \left( j_i\right) _{*}=$ $\mathrm{codim}_{X^i}\left(
X_i\right) $):
\[
A_G^{*}\left( X_2\right) \stackrel{\left( j_2\right) _{*}}{\longrightarrow }%
A_G^{*}\left( X^2\right) \stackrel{i_2^{*}}{\longrightarrow }A_G^{*}\left(
X_1\right) \rightarrow 0
\]
\begin{equation}
A_G^{*}\left( X_3\right) \stackrel{\left( j_3\right) _{*}}{\longrightarrow }%
A_G^{*}\left( X^3\right) \stackrel{i_3^{*}}{\longrightarrow }A_G^{*}\left(
X^2\right) \rightarrow 0 \label{uno}
\end{equation}
\[
\vdots
\]
\[
A_G^{*}\left( X_n\right) \stackrel{\left( j_n\right) _{*}}{\longrightarrow }%
A_G^{*}\left( X=X^n\right) \stackrel{i_n^{*}}{\longrightarrow }A_G^{*}\left(
X^{n-1}=X\setminus X_n\right) \rightarrow 0.
\]
Note that if $X$ is smooth, each graded group above is indeed a graded ring.
This will be our case.
Let
\begin{eqnarray*}
U &\doteq &\left\{ A\in \mathrm{sl}_3\backslash \left\{ 0\right\} \mid \text{
}A\text{ has distinct eigenvalues}\right\} \stackunder{\text{open}}{\subset }%
\mathrm{sl}_3\backslash \left\{ 0\right\} , \\
\text{ }Z_{1,1} &\doteq &\left\{ A\in \mathrm{sl}_3\backslash \left\{
0\right\} \mid \text{ }A\text{ has Jordan form }\left( \text{%
\begin{tabular}{lll}
$\lambda $ & $0$ & $0$ \\
$1$ & $\lambda $ & $0$ \\
$0$ & $0$ & $-2\lambda $%
\end{tabular}
}\right) \text{, }\lambda \in \mathbf{C}^{*}\right\} , \\
Z_{1,0} &\doteq &\left\{ A\in \mathrm{sl}_3\backslash \left\{ 0\right\} \mid
\text{ }A\text{ has Jordan form }\left( \text{%
\begin{tabular}{lll}
$\lambda $ & $0$ & $0$ \\
$0$ & $\lambda $ & $0$ \\
$0$ & $0$ & $-2\lambda $%
\end{tabular}
}\right) \text{, }\lambda \in \mathbf{C}^{*}\right\} , \\
Z_1 &\doteq &Z_{1,1}\cup Z_{1,0}\text{ ,} \\
Z_{0,1} &\doteq &\left\{ PGL_3\text{-orbit of }\left( \text{%
\begin{tabular}{lll}
$0$ & $0$ & $0$ \\
$1$ & $0$ & $0$ \\
$0$ & $1$ & $0$%
\end{tabular}
}\right) \right\} , \\
Z_{0,0} &\doteq &\left\{ PGL_3\text{-orbit of }\left( \text{%
\begin{tabular}{lll}
$0$ & $0$ & $0$ \\
$1$ & $0$ & $0$ \\
$0$ & $0$ & $0$%
\end{tabular}
}\right) \right\} , \\
Z_0 &\doteq &Z_{0.1}\cup Z_{0,0}\text{ ,} \\
&& \\
&&
\end{eqnarray*}
(note that $Z_1\cup Z_0=\mathrm{sl}_3\backslash (U\cup \left\{ 0\right\} )$%
). Then
\begin{equation}
\left\{ U,Z_{1,1},Z_{1,0},Z_{0,1},Z_{0,0},\left\{ 0\right\} \right\}
\label{due}
\end{equation}
is a finite $PGL_3$-stratification of $\mathrm{sl}_3$. In this case the
first associated exact sequence of (\ref{uno}) is
\begin{equation}
A_{PGL_3}^{*}\left( Z_{1,1}\right) \stackrel{\left( j_{1,1}\right) _{*}}{%
\longrightarrow }A_{PGL_3}^{*}\left( \mathrm{sl}_3\backslash (Z_{1,0}\cup
Z_0\cup \left\{ 0\right\} )\right) \stackrel{i_{1,1}^{*}}{\longrightarrow }%
A_{PGL_3}^{*}\left( U\right) \rightarrow 0 \label{tre}
\end{equation}
where $i_{1,1}:U\hookrightarrow \mathrm{sl}_3\backslash (Z_{1,0}\cup Z_0\cup
\left\{ 0\right\} )$ and $j_{1,1}:Z_{1,1}\hookrightarrow \mathrm{sl}%
_3\backslash (Z_{1,0}\cup Z_0\cup \left\{ 0\right\} )$ are the natural
immersions (open and closed, respectively).
To begin with, let us study $A_{PGL_3}^{*}\left( U\right) $.
\subsection{Generators coming from the open subset $U\subset \mathrm{sl}_3$}
Let $T$ be the maximal torus of $PGL_3$ and $\Gamma _3\doteq N_{PGL_3}\left(
T\right) =S_3\ltimes T$ its normalizer in $PGL_3$. Let $S_3\hookrightarrow
PGL_3:\sigma \mapsto \underline{\sigma }$ be the obvious inclusion (which
identifies permutations with permutation matrices). $\Gamma _3$ acts on the
subscheme $Diag_{\mathrm{sl}_3}^{*}\subset \mathrm{sl}_3\backslash \left\{
0\right\} $ of diagonal matrices with distinct eigenvalues, through $%
S_3\hookrightarrow PGL_3$
\[
\left( \sigma ,\left[ \underline{t}\right] \right) \cdot diag\left( \lambda
_1,\lambda _2,\lambda _3\right) \doteq \underline{\sigma }\cdot diag\left(
\lambda _1,\lambda _2,\lambda _3\right) \cdot \underline{\sigma }^{-1}
\]
and we have\footnote{%
This Proposition holds (with the same proof given below) for any $PGL_n$.}:
\begin{proposition}
\label{p3.1}The composition of natural maps
\[
A_{PGL_3}^{*}\left( U\right) \longrightarrow A_{\Gamma _3}^{*}\left(
U\right) \longrightarrow A_{\Gamma _3}^{*}\left( Diag_{\mathrm{sl}%
_3}^{*}\right)
\]
is a ring isomorphism.
\end{proposition}
\TeXButton{Proof}{\proof} Let $T$ act by multiplication on the right of $%
PGL_3$ and $\frac{PGL_3}T$ be the corresponding quotient. $S_3$ acts on the
left of $\frac{PGL_3}T$ via $\sigma \cdot \left[ g\right] \doteq \left[
g\sigma ^{-1}\right] ,$ $g\in PGL_3$, and on $Diag_{\mathrm{sl}_3}^{*}$ as
above. If we let $PGL_3$ act on $Diag_{\mathrm{sl}_3}^{*}\times \frac{PGL_3}%
T $ by left multiplication on $\frac{PGL_3}T$ only, there is a $PGL_3$-
equivariant isomorphism
\begin{eqnarray*}
U &\simeq &\left( Diag_{\mathrm{sl}_3}^{*}\times \frac{PGL_3}T\right)
\diagup S_3 \\
A &\longmapsto &\left[ \Delta ,\left[ g\right] _T\right] _{S_3}
\end{eqnarray*}
where $g^{-1}Ag=\Delta $ (diagonal).
Since $S_3$ acts freely on $Diag_{\mathrm{sl}_3}^{*}\times \frac{PGL_3}T$,
from Lemma \ref{change}, we get
\[
A_{PGL_3}^{*}\left( U\right) \simeq A_{PGL_3\times S_3}^{*}\left( Diag_{%
\mathrm{sl}_3}^{*}\times \frac{PGL_3}T\right) .
\]
Now, if $W$ is a free open subset of a $PGL_3\times S_3$-representation with
complement of sufficiently high codimension, we let $\Gamma _3$ act on $W$
via the inclusion
\[
\left( i,\pi \right) :\Gamma _3\hookrightarrow PGL_3\times S_3:\left( \sigma
,\left[ \underline{t}\right] \right) \longmapsto \left( \left[ \underline{t}%
\right] \underline{\sigma },\sigma \right)
\]
$i$ being the natural inclusion $\Gamma _3\hookrightarrow PGL_3$. Then the
morphisms
\begin{eqnarray*}
&&\frac{W\times Diag_{\mathrm{sl}_3}^{*}\times \frac{PGL_3}T}{PGL_3\times S_3%
}\stackrel{\psi }{\stackunder{\varphi }{\leftrightarrows }}\frac{W\times
Diag_{\mathrm{sl}_3}^{*}}{\Gamma _3} \\
\varphi &:&\left[ w,\Delta ,\left[ g\right] _T\right] _{PGL_3\times
S_3}\longmapsto \left[ w\cdot \left( g,1\right) ,\Delta \right] _{\Gamma _3}
\\
\psi &:&\left[ w,\Delta \right] _{\Gamma _3}\longmapsto \left[ w,\Delta
,\left[ 1\right] _T\right] _{PGL_3\times S_3}
\end{eqnarray*}
are mutually inverse and we conclude. \TeXButton{End Proof}{\endproof}
\begin{lemma}
\label{invariants}If $T$ denotes the maximal torus of $PGL_3$ and $A_T^{*}\ $%
is viewed as a subring of $A_{T_{GL_3}}^{*}=\mathbf{Z}\left[
x_1,x_2,x_3\right] $, then the Weyl group-invariant subring $\left(
A_T^{*}\right) ^{S_3}$ is generated by
\[
\gamma _2=s_1^2-3s_2
\]
\[
\gamma _3=2s_1^3-9s_1s_2+27s_3
\]
\[
\gamma _6=\Delta \equiv (x_1-x_2)^2(x_1-x_3)^2(x_2-x_3)^2
\]
where $s_i$ denotes the $i$-th elementary symmetric function on the $x_j$'s
and $\Delta $ is the discriminant.
\end{lemma}
\TeXButton{Proof}{\proof} We have $T=T_{PGL_3}\simeq T_{GL_3}/\mathbf{G}_m$,
where $T_{GL_3}=\left( \mathbf{G}_m\right) ^3$ and $G_m\hookrightarrow
T_{GL_3}$ diagonally. Therefore
\[
A_T^{*}=Sym_{\mathbf{Z}}\left( \widehat{T}\right) \subset
A_{T_{GL_3}}^{*}=Sym_{\mathbf{Z}}\left( \widehat{T_{GL_3}}\right) =\mathbf{Z}%
\left[ x_1,x_2,x_3\right]
\]
is the subring of polynomials $f\left( x_1,x_2,x_3\right) $ such that
\[
f\left( x_1+t,x_2+t,x_3+t\right) =f\left( x_1,x_2,x_3\right) .
\]
Then
\[
\left( A_T^{*}\right) ^{S_3}=\left\{ f\in \mathbf{Z}\left[
x_1,x_2,x_3\right] ^{S_3}\mid f\left( x_1+t,x_2+t,x_3+t\right) =f\left(
x_1,x_2,x_3\right) \right\} \equiv
\]
\[
\equiv \left( \mathbf{Z}\left[ s_1,s_2,s_3\right] \right) ^{inv}
\]
where $S_3$ permutes the $x_i$'s. Now, if for any polynomial $f\in \mathbf{Z}%
\left[ x_1,x_2,x_3\right] $ we let $f^t=f\left( x_1+t,x_2+t,x_3+t\right) $,
we get
\begin{eqnarray}
s_1^t &=&s_1+3t \label{s1} \\
s_2^t &=&s_2+2s_1t+3t^2 \label{s2} \\
s_3^t &=&s_3+s_2t+s_1t^2+t^3 \label{s3}
\end{eqnarray}
and it is then easy to verify that $\gamma _2,\gamma _3$ and $\gamma _6$ are
indeed in $\left( A_T^{*}\right) ^{S_3}$.
Now, let $\varphi \in \left( A_T^{*}\right) ^{S_3}$. We first claim that
there exists $n_\varphi \geq 0$ such that $3^{n_\varphi }\varphi \in \mathbf{%
Z}\left[ \gamma _2,\gamma _3,\gamma _6\right] $. By definition of $\gamma _2$
and $\gamma _3$, we have
\[
\left( A_T^{*}\right) ^{S_3}\left[ \frac 13\right] \equiv \left( \mathbf{Z}%
\left[ \frac 13\right] \left[ s_1,s_2,s_3\right] \right) ^{inv}=\left(
\mathbf{Z}\left[ \frac 13\right] \left[ s_1,\gamma _2,\gamma _3\right]
\right) ^{inv}\text{.}
\]
If
\[
P\left( s_1,\gamma _2,\gamma _3\right) =P_0\left( \gamma _2,\gamma _3\right)
+P_1\left( \gamma _2,\gamma _3\right) s_1+\cdots +P_m\left( \gamma _2,\gamma
_3\right) s_1^m
\]
is in $\in \left( \mathbf{Z}\left[ \frac 13\right] \left[ s_1,\gamma
_2,\gamma _3\right] \right) ^{inv}$, using (\ref{s1}) and $\gamma
_2^t=\gamma _2$, $\gamma _3^t=\gamma _3$, we easily get, by induction on $m$%
, $P_i=0,$ $\forall i\geq 1$, i.e.
\[
\left( A_T^{*}\right) ^{S_3}\left[ \frac 13\right] =\mathbf{Z}\left[ \frac
13\right] \left[ \gamma _2,\gamma _3\right]
\]
as claimed.
To prove that indeed $\varphi \in \mathbf{Z}\left[ \gamma _2,\gamma
_3,\gamma _6\right] $, we use induction on $n_\varphi $.
Suppose\footnote{%
Note that we allow an explicit dependence of $p$ on $\gamma _6$!} $3\varphi
=p\left( \gamma _2,\gamma _3,\gamma _6\right) $, for some polynomial $p$.
Expanding $p$ in powers of $\gamma _6$, we get
\[
3\varphi =p_0\left( \gamma _2,\gamma _3\right) +p_1\left( \gamma _2,\gamma
_3\right) \gamma _6+\cdots
\]
and reducing $\limfunc{mod}3$%
\[
0\equiv p_0\left( s_1^2,-s_1^3\right) +p_1\left( s_1^2,-s_1^3\right) \gamma
_6+\cdots \qquad \left( \limfunc{mod}3\right) .
\]
But $s_1$ and $\gamma _6=\Delta $ are algebraically independent (over $%
\mathbf{Z}/3$), so $p_i\left( s_1^2,-s_1^3\right) \equiv 0$ $\left( \limfunc{%
mod}3\right) $, $\forall i$, i.e.
\[
p_i\left( s_1^2,-s_1^3\right) =\left( \left( s_1^2\right) ^3-\left(
s_1^3\right) ^2\right) \cdot q_i\left( s_1^2,s_1^3\right) \qquad \left(
\limfunc{mod}3\right)
\]
then
\[
p_i\left( \gamma _2,\gamma _3\right) =\left( \gamma _2^3-\gamma _3^2\right)
q_i\left( \gamma _2,-\gamma _3\right) +3r_i\left( \gamma _2,\gamma _3\right)
\]
for each $i$. Thus
\[
3\varphi =3r\left( \gamma _2,\gamma _3,\gamma _6\right) +\left( \gamma
_2^3-\gamma _3^2\right) q\left( \gamma _2,\gamma _3,\gamma _6\right) .
\]
with an obvious notation. Straightforward computations yield
\[
\left( \gamma _2^3-\gamma _3^2\right) =-3\left( \gamma _2^3-9\gamma
_6\right) \text{,}
\]
and the case $n_\varphi =1$ is settled. The inductive step follows easily
from the fact that we included a possible dependence of $p$ from $\gamma _6$
in the above argument. \TeXButton{End Proof}{\endproof}
\begin{remark}
\label{twistaction}Note that there is a (non canonical) isomorphism
\[
T\longrightarrow \left( \mathbf{G}_m\right) ^2
\]
\[
\left[ t_1,t_2,t_3\right] \longmapsto \left( t_1/t_3,t_2/t_3\right)
\]
so that $A_T^{*}\simeq \mathbf{Z}\left[ x,y\right] $, with action of the
Weyl group given by
\[
\left( 12\right) x=y,\text{ }\left( 12\right) y=x
\]
\begin{equation}
\left( 123\right) x=-y,\text{ }\left( 123\right) y=x-y. \label{twistaction1}
\end{equation}
Under this isomorphism, with the same notations as in lemma \ref{invariants}%
, we have
\[
\gamma _2=\left( x+y\right) ^2-3xy,
\]
\begin{equation}
\gamma _3=-9\left( x+y\right) xy+2\left( x+y\right) ^3 \label{twovariables}
\end{equation}
\[
\gamma _6=\left( x+y\right) ^2x^2y^2-4x^3y^3.
\]
Moreover, there is an isomorphism of $T$ with $T_{SL_3}$, the maximal torus
of $SL_3$%
\begin{equation}
T\rightarrow T_{SL_3}:\left[ t_1,t_2,t_3\right] \longmapsto \left(
t_2/t_3,t_3/t_1,t_1/t_2\right) \label{twistaction2}
\end{equation}
and an induced isomorphism $A_T^{*}\simeq A_{T_{SL_3}}^{*}=\mathbf{Z}\left[
u_1,u_2,u_3\right] \diagup \left( u_1+u_2+u_3\right) $. The Weyl groups are
isomorphic to $S_3$ in both cases but the isomorphism above on Chow rings is
not $S_3$-equivariant, only $A_3$-equivariant. Rather, the action of $S_3$
on $A_{T_{SL_3}}^{*}$ inherited from the Weyl group action on $A_T^{*}$ via
this isomorphism, is given by
\begin{eqnarray*}
\left( 12\right) u_1=-u_2,\text{ }\left( 12\right) u_2=-u_1,\,\left(
12\right) u_3=-u_3 \\
\left( 123\right) u_1=u_3,\text{ }\left( 123\right) u_2=u_1,\,\left(
123\right) u_3=u_1.
\end{eqnarray*}
\end{remark}
\begin{corollary}
\label{hsurj}The canonical morphism $h:A_{\Gamma _3}^{*}\rightarrow \left(
A_T^{*}\right) ^{S_3}$ is surjective.
\end{corollary}
\TeXButton{Proof}{\proof} Let $\phi :A_{PGL_3}^{*}\rightarrow A_{\Gamma
_3}^{*}$ be the restriction morphism, $E$ the standard representation of $%
GL_3$ and $Sym^3E$ be the $PGL_3$-representation:
\[
\left[ g\right] \cdot \left( v\cdot _1v_2\cdot v_3\right) \doteq \left( \det
g^{-1}\right) \left( gv_1\cdot gv_2\cdot gv_3\right) .
\]
It is not difficult to verify that
\begin{eqnarray*}
h\circ \phi \left( c_2\left( \mathrm{sl}_3\right) \right) &=&-2\gamma _2 \\
\text{ }h\circ \phi \left( c_2\left( Sym^3E\right) \right) &=&-5\gamma _2
\end{eqnarray*}
\begin{eqnarray*}
\text{ }h\circ \phi \left( c_3\left( Sym^3E\right) \right) &=&\gamma _3 \\
\text{ }h\circ \phi \left( c_6\left( \mathrm{sl}_3\right) \right) &=&\gamma
_6
\end{eqnarray*}
and the Corollary follows from Lemma \ref{invariants}. \TeXButton{End Proof}
{\endproof}
Now consider the subgroup $A_3\ltimes T\hookrightarrow \Gamma _3=S_3\ltimes
T $; there is a transfer morphism (see (\ref{tsf}), Section 2)
\[
\mathrm{tsf}=\mathrm{tsf}_{A_3\ltimes T}^{\Gamma _3}\left( Diag_{\mathrm{sl}%
_3}^{*}\right) :A_{A_3\ltimes T}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right)
\rightarrow A_{\Gamma _3}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right)
\]
and a restriction morphism:
\[
\mathrm{res}=\mathrm{res}_{A_3\ltimes T}^{\Gamma _3}\left( Diag_{\mathrm{sl}%
_3}^{*}\right) :A_{\Gamma _3}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right)
\rightarrow \left( A_{A_3\ltimes T}^{*}\left( Diag_{\mathrm{sl}%
_3}^{*}\right) \right) ^{C_2}.
\]
\begin{lemma}
\label{transfertrick}\emph{(transfer-trick)} $\mathrm{res}$ induces an
isomorphism
\[
_3A_{\Gamma _3}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right) \rightarrow \text{
}_3\left( A_{A_3\ltimes T}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right) \right)
^{C_2}
\]
with inverse $(-\mathrm{tsf})$.
\end{lemma}
\TeXButton{Proof}{\proof} By projection formula, $\mathrm{tsf}\circ \mathrm{%
res}=2$; so if $\xi $ is $3$-torsion, we have $\mathrm{tsf}\circ \mathrm{res}%
\left( \xi \right) =-\xi $. On the other hand, if $C_2=\left\{ 1,\varepsilon
\right\} $, we have $\mathrm{res}\circ \mathrm{tsf}\left( \eta \right) =\eta
+\eta ^\varepsilon $; so, if $\eta $ is $C_2$-invariant and $3$-torsion, we
have $\mathrm{res}\circ \mathrm{tsf}\left( \eta \right) =-\eta $ and
conclude. \TeXButton{End Proof}{\endproof}
The isomorphism (\ref{twistaction2}) of Rmk. \ref{twistaction} induces an
isomorphism
\[
A_3\ltimes T\simeq A_3\ltimes T_{SL_3}
\]
and hence an isomorphism
\begin{equation}
A_{A_3\ltimes T}^{*}\simeq A_{A_3\ltimes T_{SL_3}}^{*}. \label{bingo}
\end{equation}
We will consider the $C_2$-action on $A_{A_3\ltimes T_{SL_3}}^{*}$ induced
by the canonical action on $A_{A_3\ltimes T}^{*}$ via this isomorphism. As
already in Rmk. \ref{twistaction}, we warn the reader that this is not the
canonical action induced by the inclusion $A_3\ltimes
T_{SL_3}\hookrightarrow N_{SL_3}\left( T_{SL_3}\right) $.
If $A_{A_3}^{*}=\mathbf{Z}\left[ \alpha \right] /(3\alpha )$ \footnote{%
We use that $A_3\simeq \mu _3$, which is true over any algebraically closed
field of characteristic $\neq 3$. Note that in characteristic $3$, it is no
longer true that $A_{A_3}^{*}\simeq \mathbf{Z}\left[ \alpha \right] /\left(
3\alpha \right) $.}, we still denote by $\alpha $ the image of $\alpha $ in $%
A_{A_3\ltimes T_{SL_3}}^{*}$ via the pullback induced by the projection $%
A_3\ltimes T_{SL_3}\rightarrow A_3$. We also recall the isomorphism $%
A_{T_{SL_3}}^{*}\simeq \mathbf{Z}\left[ u_1,u_2,u_3\right] \diagup \left(
u_1+u_2+u_3\right) $.
Then, if $W\simeq \mathbf{C}^3$ denotes the $A_3\ltimes T_{SL_3}$%
-representation
\begin{equation}
\text{ }\left( \sigma ,\left( \underline{s}\right) \right) \cdot \left(
\underline{x}\right) \doteq \left( s_1x_{\sigma ^{-1}\left( 1\right)
},s_2x_{\sigma ^{-1}\left( 2\right) },s_3x_{\sigma ^{-1}\left( 3\right)
}\right) \text{,} \label{vudoppio}
\end{equation}
we have the following basic result
\begin{proposition}
\label{basico}The ring $A_{A_3\ltimes T_{SL_3}}^{*}$ is generated by
\[
\left\{ \alpha ,\text{ }c_2\left( W\right) ,\text{ }c_3\left( W\right) ,%
\text{ }\theta \doteq \mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}\left(
u_2^2u_3\right) \right\} .
\]
\end{proposition}
\TeXButton{Proof}{\proof}Throughout the proof we identify $A_3\ltimes T$
with $A_3\ltimes T_{SL_3}$ (Remark \ref{twistaction}). $A_3\ltimes T_{SL_3}$
acts on $\mathbf{P}\left( W\right) $ with a dense orbit $U\doteq D_{+}\left(
x_1x_2x_3\right) $ with stabilizer isomorphic to $A_3\times \mathbf{\mu }_3$%
. If $j_2:Y_2\hookrightarrow \mathbf{P}\left( W\right) $ denotes the
(closed) orbit of $\left[ 1,0,0\right] \in \mathbf{P}\left( W\right) $ and
\[
Y_1\doteq \mathbf{P}\left( W\right) \backslash U\cup Y_2\stackrel{j_1}{%
\stackunder{^{closed}}{\hookrightarrow }}\mathbf{P}\left( W\right)
\backslash Y_2,
\]
the orbit of $\left[ 1,1,0\right] \in \mathbf{P}\left( W\right) $, then $%
\left\{ U,Y_1,Y_2\right\} $ is a finite $A_3\ltimes T_{SL_3}$-stratification
of $\mathbf{P}\left( W\right) $ and the exact sequences (\ref{uno}) are
\begin{equation}
A_{\mathbf{G}_m}^{*}\simeq A_{A_3\ltimes T_{SL_3}}^{*}\left( Y_1\right)
\stackrel{\left( j_1\right) _{*}}{\longrightarrow }A_{A_3\ltimes
T_{SL_3}}^{*}\left( \mathbf{P}\left( W\right) \backslash Y_2\right)
\stackrel{i^{*}}{\longrightarrow }A_{A_3\ltimes T_{SL_3}}^{*}\left( U\right)
\simeq A_{A_3\times \mathbf{\mu }_3}^{*}\rightarrow 0 \label{prima}
\end{equation}
\begin{equation}
A_{T_{SL_3}}^{*}\simeq A_{A_3\ltimes T_{SL_3}}^{*}\left( Y_2\right)
\stackrel{\left( j_2\right) _{*}}{\longrightarrow }A_{A_3\ltimes
T_{SL_3}}^{*}\left( \mathbf{P}\left( W\right) \right) \stackrel{i_2^{*}}{%
\longrightarrow }A_{A_3\ltimes T_{SL_3}}^{*}\left( \mathbf{P}\left( W\right)
\backslash Y_2\right) \rightarrow 0. \label{seconda}
\end{equation}
where we used Prop. \ref{homogeneous} together with the fact that $Y_1$
(resp. $Y_2$, $U$) has stabilizer isomorphic to $\mathbf{G}_m$ (resp. $%
T_{SL_3}$, $A_3\times \mathbf{\mu }_3$). By (\cite{Fu}, Th. 3.3 (b)), we
have ($c_1\left( W\right) =0$) :
\[
A_{A_3\ltimes T_{SL_3}}^{*}\left( \mathbf{P}\left( W\right) \right) \simeq
A_{A_3\ltimes T_{SL_3}}^{*}\left[ \ell \right] \diagup \left( \ell
^3+c_2\left( W\right) \ell +c_3\left( W\right) \right)
\]
where $\ell =c_1\left( \mathcal{O}_{\mathbf{P}\left( W\right) }\left(
1\right) \right) $. Moreover, the canonical K\"unneth morphism
\[
A_{A_3}^{*}\otimes A_{\mathbf{\mu }_3}^{*}\simeq \mathbf{Z}\left[ \alpha
\right] /\left( 3\alpha \right) \otimes \mathbf{Z}\left[ \beta \right]
/\left( 3\beta \right) \rightarrow A_{A_3\times \mathbf{\mu }_3}^{*}
\]
is an isomorphism (e.g. \cite{To2}, \S\ 6). It is not difficult to show that
\[
i^{*}\left( \ell \right) =-\beta \text{, }i^{*}\left( \alpha \right) =\alpha
\text{, }j_1^{*}\left( \ell \right) =-u
\]
(where $A_{\mathbf{G}_m}^{*}=\mathbf{Z}\left[ u\right] $ and with the usual
abuse of notation, we write $\ell $ for $i_2^{*}\left( \ell \right) $ and $%
\alpha $ for its pullback to $A_{A_3\ltimes T_{SL_3}}^{*}\left( \mathbf{P}%
\left( W\right) \backslash Y_2\right) $). So we can conclude the analysis of
(\ref{prima}), by computing $\left( j_1\right) _{*}\left( 1\right) =\left[
Y_1\right] $. $Y_1$ is the zero scheme of the $A_3\ltimes T_{SL_3}$%
-invariant regular section $x_1x_2x_3\in \Gamma \left( \mathcal{O}\left(
3\right) ,\mathbf{P}\left( W\right) \backslash Y_2\right) $, hence (\cite{Fu}%
, p. 61), $\left[ Y_1\right] =3\ell $ so that $A_{A_3\ltimes
T_{SL_3}}^{*}\left( \mathbf{P}\left( W\right) \backslash Y_2\right) $ is
generated by $\left\{ \alpha ,\ell \right\} $.
Now let us turn our attention to (\ref{seconda}). It is easy to verify that,
with the usual abuse of notation, $i_2^{*}\left( \alpha \right) =\alpha $
and $i_2^{*}\left( \ell \right) =\ell $, so we are left to find generators
of $A_{A_3\ltimes T_{SL_3}}^{*}\left( Y_2\right) \simeq A_{T_{SL_3}}^{*}$ as
an $A_{A_3\ltimes T_{SL_3}}^{*}\left( \mathbf{P}\left( W\right) \right) $%
-module.
First of all, we have $j_2{}^{*}\left( \ell \right) =u_1$ \footnote{%
Of course this relation depends on the choice of the isomorphism
\[
\mathbf{Z}\left[ u_1,u_2,u_3\right] \diagup \left( u_1+u_2+u_3\right)
=A_{T_{SL_3}}^{*}\simeq A_{A_3\ltimes T_{SL_3}}^{*}\left( Y_2\right)
\]
which in its turn depends essentially on the choice of a point
\[
p\in Y_2=\left\{ \left[ 1,0,0\right] ,\left[ 0,1,0\right] ,\left[
0,0,1\right] \right\} .
\]
The choice we are making here is $p=\left[ 1,0,0\right] $.}. Therefore, by
projection formula and the relation $u_1+u_2+u_3=0$, we see that $%
A_{T_{SL_3}}^{*}\simeq A_{A_3\ltimes T_{SL_3}}^{*}\left( Y_2\right) $ is
generated by
\begin{equation}
\left\{ 1,u_2^n\mid n>0\right\} \label{first step}
\end{equation}
as an $A_{A_3\ltimes T_{SL_3}}^{*}\left( \mathbf{P}\left( W\right) \right) $%
-module. But
\[
j_2{}^{*}\left( c_2\left( W\right) \right) =u_1u_2+u_2u_3+u_3u_1=-\left(
u_1^2+u_2^2+u_1u_2\right)
\]
so that, by induction on $n$, $\left( j_2\right) _{*}\left( u_2^n\right) $, $%
n>1$, belongs to the submodule generated by $\left( j_2\right) _{*}\left(
1\right) $ and $\left( j_2\right) _{*}\left( u_2\right) $ (e.g.
\begin{eqnarray*}
\left( j_2\right) _{*}\left( u_2^2\right) &=&\left( j_2\right) _{*}\left(
j_2^{*}\left( -c_2\left( W\right) \right) \right) +\left( j_2\right)
_{*}\left( j_2^{*}\left( -\ell ^2\right) \right) -\left( j_2\right)
_{*}\left( j_2^{*}\left( \ell \right) \cdot u_2\right) = \\
&=&-c_2\left( W\right) \cdot \left( j_2\right) _{*}\left( 1\right) -\ell
^2\cdot \left( j_2\right) _{*}\left( 1\right) -\ell \cdot \left( j_2\right)
_{*}\left( u_2\right)
\end{eqnarray*}
and similarly for higher powers of $u_2$). Thus, the ideal
\[
im\left( j_2\right) _{*}\subset A_{A_3\ltimes T_{SL_3}}^{*}\left( \mathbf{P}%
\left( W\right) \right)
\]
is actually generated by $\left( j_2\right) _{*}\left( 1\right) $ and $%
\left( j_2\right) _{*}\left( u_2\right) $.
Let us first compute $\left( j_2\right) _{*}\left( 1\right) $ using a
transfer argument (Section 2). Consider the $A_3\ltimes T_{SL_3}$-
equivariant commutative diagram
\[\begin{tabular}{ll}
$Y_2\stackrel{h_2}{\longrightarrow }$ & $\mathbf{P}\left( W\right) \times A_3
$ \\
$\quad _{j_2}\searrow $ & $\quad \downarrow ^{pr_1}$ \\
& $\mathbf{P}\left( W\right) $%
\end{tabular}
\]
where
\begin{eqnarray*}
h_2\left( \left[ 1,0,0\right] \right) =\left( \left[ 1,0,0\right] ,1\right) ,%
\text{ }h_2\left( \left[ 0,1,0\right] \right) =\left( \left[ 0,1,0\right]
,\sigma \right) ,\text{ }h_2\left( \left[ 0,0,1\right] \right) =\left(
\left[ 0,0,1\right] ,\sigma ^2\right)
\end{eqnarray*}
with $\sigma =\left( 123\right) $. Using the canonical isomorphism
\[
A_{A_3\ltimes T_{SL_3}}^{*}\left( \mathbf{P}(W)\times A_3\right) \simeq
A_T^{*}\left( \mathbf{P}(W)\right)
\]
we see that
\begin{equation}
\left( j_2\right) _{*}\left( 1\right) =\left( pr_1\right) _{*}\circ \left(
h_2\right) _{*}\left( 1\right) =\mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes
T_{SL_3}}\left( \mathbf{P}(W)\right) \left( \left[ \left\{ \left[
1,0,0\right] \right\} \right] \right) . \label{alfa}
\end{equation}
But $\left[ 1,0,0\right] =Z\left( x_2\right) \cap Z\left( x_3\right) $,
where the section $x_i$, $i=2,3$ are $T_{SL_3}$-semi-invariant (\cite{SGA3.I}%
, Expos\'e VI$_{\mathrm{B}},$ p. 406) so that if we consider the $T_{SL_3}$%
-equivariant line bundles $L_i\rightarrow Spec\mathbf{C}$ associated to the
representations
\[
\left( \underline{t}\right) x=t_ix\text{,\quad \quad }i=2,3\text{,}
\]
we have induced $T_{SL_3}$-invariant regular sections $\widetilde{x_i}\in
\Gamma \left( P(W),\mathcal{O}\left( 1\right) \otimes p^{*}\left( L_i^{\vee
}\right) \right) $ \footnote{%
Note that $p^{*}\left( L_i^{\vee }\right) $ is trivial but not $T_{SL_3}$%
-equivariantly trivial.} with, obviously, $Z\left( \widetilde{x_i}\right)
=Z\left( x_i\right) $. Then
\begin{equation}
\left[ \left\{ \left[ 1,0,0\right] \right\} \right] =\left( \ell -u_2\right)
\left( \ell -u_3\right) =\ell ^2+\ell u_1+u_2u_3 \label{beta}
\end{equation}
in $A_{T_{SL_3}}^{*}\left( \mathbf{P}(W)\right) $. Since $\ell =\mathrm{res}%
_{A_3\ltimes T_{SL_3}}^{T_{SL_3}}\left( \mathbf{P}(W)\right) \left( \ell
\right) $ and the diagram
\[
\begin{tabular}{ccc}
$\qquad \qquad A_{T_{SL_3}}^{*}$ & $\longrightarrow $ & $A_{T_{SL_3}}^{*}%
\left( \mathbf{P}(W)\right) $ \\
$^{\mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}}\downarrow $ & & $\qquad
\quad \downarrow ^{\mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}\left(
\mathbf{P}(W)\right) }$ \\
$\qquad \qquad A_{A_3\ltimes T_{SL_3}}^{*}$ & $\longrightarrow $ & $%
A_{A_3\ltimes T_{SL_3}}^{*}\left( \mathbf{P}(W)\right) $%
\end{tabular}
\]
is commutative, we have
\begin{eqnarray}
\mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}\left( \mathbf{P}(W)\right)
\left( \ell ^2\right) &=&3\ell ^2 \label{gamma} \\
\mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}\left( \mathbf{P}(W)\right)
\left( \ell u_1\right) &=&\ell \cdot \mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes
T_{SL_3}}\left( u_1\right) . \label{gamma2}
\end{eqnarray}
Now we claim $\mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}\left( u_1\right)
=0$, $i=1,2,3$. In fact, let $\pi :A_{A_3\ltimes T_{SL_3}}^{*}\rightarrow A_3
$ be the projection and $\rho :A_3\hookrightarrow A_{A_3\ltimes T_{SL_3}}^{*}
$ its right inverse. Since $i_2^{*}$ in (\ref{seconda}) is an isomorphism in
degree $1$ and $A_{A_3\ltimes T_{SL_3}}^{*}\left( \mathbf{P}\left( W\right)
\backslash Y_2\right) $ is generated by $\alpha $ and $\ell $, $\mathrm{tsf}%
_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}\left( u_i\right) =n_i\pi ^{*}\alpha $ for
some integer $n_i$ (in fact
\[
\mathrm{res}_{A_3\ltimes T_{SL_3}}^{T_{SL_3}}\circ \mathrm{tsf}%
_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}\left( u_i\right) =u_1+u_2+u_3=0
\]
thus $\mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}\left( u_i\right) $ is $3$%
-torsion). Since
\[
\rho ^{*}\circ \mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}\equiv \mathrm{%
res}_{A_3}^{A_3\ltimes T_{SL_3}}\circ \mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes
T_{SL_3}}=0,
\]
we get
\[
n_i\rho ^{*}\pi ^{*}\left( \alpha \right) =n_i\alpha =0
\]
in $A_{A_3}^{*}$ and the claim follows.
Since $\left( j_2\right) _{*}\left( u_2\right) $ has degree $3$, from (\ref
{seconda}) and the computations we have just done (in particular (\ref{alfa}%
), (\ref{beta}), (\ref{gamma}) and (\ref{gamma2})), we know that the ring $%
A_{A_3\ltimes T_{SL_3}}^{*}\left( \mathbf{P}\left( W\right) \right) $ is
generated up to degree $2$ (included) by
\[
\left\{ \alpha ,\ell ,\mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}\left(
u_2u_3\right) \right\} .
\]
We will show that:
\texttt{CLAIM}. $A_{A_3\ltimes T_{SL_3}}^{*}\left( \mathbf{P}\left( W\right)
\right) $ is generated up to degree $2$ (included) by
\[
\left\{ \alpha ,\ell ,c_2\left( W\right) \right\} .
\]
\underline{Proof of Claim}. We write
\[
\eta _{\mid T_{SL_3}}\equiv \mathrm{res}_{A_3\ltimes
T_{SL_3}}^{T_{SL_3}}\left( \eta \right) ,
\]
for any $\eta \in A_{A_3\ltimes T_{SL_3}}^{*}\left( \mathbf{P}\left(
W\right) \right) $.
Observe that
\[
\mathrm{res}_{A_3\ltimes T_{SL_3}}^{T_{SL_3}}\circ \mathrm{tsf}%
_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}\left( u_2u_3\right)
=u_2u_3+u_3u_1+u_1u_2=c_2\left( W\right) _{\mid T_{SL_3}}\text{,}
\]
therefore $\mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}\left( u_2u_3\right)
-c_2\left( W\right) =\xi $, for some $3$-torsion element\footnote{%
In fact $\mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}\circ \mathrm{res}%
_{A_3\ltimes T_{SL_3}}^{T_{SL_3}}=3$.}
\[
\xi \in A_{A_3\ltimes T_{SL_3}}^2\left( \mathbf{P}\left( W\right) \right) .
\]
Since the group $A_{A_3\ltimes T_{SL_3}}^2\left( \mathbf{P}\left( W\right)
\right) $ is generated by
\[
\left\{ \alpha ^2,\ell ^2,\alpha \ell ,\mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes
T_{SL_3}}\left( u_2u_3\right) =c_2\left( W\right) +\xi \right\}
\]
we have
\begin{equation}
c_2\left( W\right) =A\left( c_2\left( W\right) +\xi \right) +B\alpha
^2+C\ell ^2+D\alpha \ell \text{.} \label{delta}
\end{equation}
Restricting to $T_{SL_3}$, we get
\[
c_2\left( W\right) _{\mid T_{SL_3}}=Ac_2\left( W\right) _{\mid
T_{SL_3}}+C\ell ^2;
\]
but from
\[
A_{T_{SL_3}}^{*}\left( \mathbf{P}(W)\right) \simeq A_{T_{SL_3}}^{*}\left[
\ell \right] \diagup \left( \ell ^3+\ell ^2c_1\left( W\right) _{\mid
T_{SL_3}}+\ell c_2\left( W\right) _{\mid T_{SL_3}}+c_3\left( W\right) _{\mid
T_{SL_3}}\right) ,
\]
we see that $c_2\left( W\right) _{\mid T_{SL_3}}$ and $\ell ^2$ are
algebraically independent, so we must have $A=1$, $C=0$. Thus (\ref{delta})
yields $\xi =B\alpha ^2+D\alpha \ell $ and this concludes the proof of Claim.%
$\ \mathbf{\Box }$
So, the other possible generators of $A_{A_3\ltimes T_{SL_3}}^{*}\left(
\mathbf{P}\left( W\right) \right) $ in degree $>2$ can only come from $%
\left( j_2\right) _{*}\left( u_2\right) $. Using the same arguments as in
the computation of $\left( j_2\right) _{*}\left( 1\right) $ above, we get
\[
\left( j_2\right) _{*}\left( u_2\right) =\mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes
T_{SL_3}}\left( \mathbf{P}(W)\right) \left( u_2\left( \ell -u_2\right)
\left( \ell -u_3\right) \right) \text{.}
\]
But, since we know that $\mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}\left(
u_i\right) =0$ $\forall i$, the only new generator is $\mathrm{tsf}%
_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}\left( u_2^2u_3\right) $.
To summarize, we have proved so far that $A_{A_3\ltimes T_{SL_3}}^{*}\left(
\mathbf{P}\left( W\right) \right) $ is generated by
\[
\left\{ \alpha ,\ell ,c_2(W),\mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes
T_{SL_3}}\left( u_2^2u_3\right) \right\} \text{.}
\]
Since
\[
A_{A_3\ltimes T_{SL_3}}^{*}\diagup \left( c_3\left( W\right) \right) \simeq
A_{A_3\ltimes T_{SL_3}}^{*}\left( W\backslash \left\{ 0\right\} \right)
\simeq A_{A_3\ltimes T_{SL_3}}^{*}\left( \mathbf{P}\left( W\right) \right)
\diagup \left( \ell \right) ,
\]
we conclude that $A_{A_3\ltimes T_{SL_3}}^{*}$ is generated by
\[
\left\{ \alpha ,c_2(W),c_3(W),\mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes
T_{SL_3}}\left( u_2^2u_3\right) \right\} .
\]
\TeXButton{End Proof}{\endproof}
Recall (\ref{bingo}) and the isomorphism
\[
_3A_{\Gamma _3}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right) \simeq \text{ }%
_3\left( A_{A_3\ltimes T}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right) \right)
^{C_2}
\]
from lemma \ref{transfertrick}. If $C_2=\left\{ 1,\varepsilon \right\} $, we
denote by $W^\varepsilon $ the $A_3\ltimes T_{SL_3}$-representation obtained
from $W$ twisting the action by $\varepsilon $. Let us also define the
element
\begin{equation}
\underline{\chi }\doteq \left( 2\mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes
T_{SL_3}}\left( u_2^2u_3\right) +3c_3\left( W\right) \right) ^2+4c_2\left(
W\right) ^3+27c_3\left( W\right) ^2\in A_{A_3\ltimes T_{SL_3}}^6 \label{chi}
\end{equation}
and denote $\mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}\left(
u_2^2u_3\right) $ simply by $\theta $.
\begin{lemma}
\label{basico2}\textrm{(i) }In $A_{A_3\ltimes T}^{*}$ we have
\[
3\underline{\chi }=3\alpha =\alpha \theta =\alpha ^3+\alpha c_2\left(
W\right) =0.
\]
\textrm{(ii)} The kernel of the restriction map $h^{\prime }:A_{A_3\ltimes
T}^{*}\rightarrow A_{A_3\ltimes T}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right)
$ is the ideal $\left( \alpha ^2\right) $.
\textrm{(iii)} In $A_{A_3\ltimes T}^{*}$, we have
\[
c_2\left( W^\varepsilon \right) =c_2\left( W\right) ,\qquad c_3\left(
W^\varepsilon \right) =-c_3\left( W\right) ,
\]
\[
\theta ^\varepsilon =\theta +3c_3\left( W\right) ,\qquad \underline{\chi }%
^\varepsilon =\underline{\chi }.
\]
\textrm{(iv)} Let
\[
q\left( c_2(W),c_3(W),\mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}\left(
u_2^2u_3\right) \right) \in \text{ }_3A_{A_3\ltimes T_{SL_3}}^{*}
\]
be a polynomial in the arguments indicated. Then there exists a polynomial
\[
\widetilde{q}=\widetilde{q}\left( c_2(W),c_3(W),\mathrm{tsf}%
_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}\left( u_2^2u_3\right) \right)
\]
such that $q=\underline{\chi }\widetilde{q}$.
\end{lemma}
\TeXButton{Proof}{\proof} (i) Since
\[
\left( 2\mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}\left( u_2^2u_3\right)
+3c_3\left( W\right) \right) _{\mid T_{SL_3}}^2=\Delta \left(
u_1,u_2,u_3\right)
\]
\begin{eqnarray*}
c_2\left( W\right) _{\mid T_{SL_3}} &=&s_2\left( u_1,u_2,u_3\right) \\
c_3\left( W\right) _{\mid T_{SL_3}} &=&s_3\left( u_1,u_2,u_3\right)
\end{eqnarray*}
in $A_{T_{SL_3}}^{*}\simeq A_T^{*}$ (where $\Delta $ is the discriminant and
$s_i$ the $i$-th elementary symmetric function), it is well known that $%
\underline{\chi }_{\mid T}=0$. Therefore $3\underline{\chi }=0$. $\alpha $
is $3$-torsion by definition and
\[
\alpha \cdot \mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}\left(
u_2^2u_3\right) =0
\]
by projection formula. Finally observe that
\[
\left( c_2\left( W\right) -\mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes
T_{SL_3}}\left( u_1u_3\right) \right) _{\mid T_{SL_3}}=0
\]
and therefore (Proposition \ref{basico}) there exist $A,B\in \mathbf{Z}$
such that
\[
c_2\left( W\right) -\mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}\left(
u_1u_2\right) =A\alpha ^2+Bc_2\left( W\right)
\]
is a $3$-torsion element in $A_{A_3\ltimes T_{SL_3}}^{*}$. Restricting to $%
T_{SL_3}$ we get $B=0$ while restricting to $A_3$ we get $A\equiv -1$ $%
\limfunc{mod}3$. Multiplying by $\alpha $, we get
\[
\alpha ^3+\alpha c_2\left( W\right) =0
\]
by projection formula.
(ii) A straightforward computation yields
\[
c_2\left( Diag_{\mathrm{sl}_3}\right) =-\alpha ^2\in A_{A_3\ltimes T}^{*}.
\]
Consider then the two localization sequences:
\begin{equation}
A_{A_3\ltimes T}^{*}\stackrel{(-\alpha ^2)\cdot }{\longrightarrow }%
A_{A_3\ltimes T}^{*}\left( Diag_{\mathrm{sl}_3}\right) \simeq A_{A_3\ltimes
T}^{*}\longrightarrow A_{A_3\ltimes T}^{*}\left( Diag_{\mathrm{sl}%
_3}\backslash \left\{ 0\right\} \right) \rightarrow 0 \label{luno}
\end{equation}
\begin{equation}
A_T^{*}\simeq A_{A_3\ltimes T}^{*}\left( Z\right) \stackrel{j_{*}}{%
\longrightarrow }A_{A_3\ltimes T}^{*}\left( Diag_{\mathrm{sl}_3}\backslash
\left\{ 0\right\} \right) \longrightarrow A_{A_3\ltimes T}^{*}\left( Diag_{%
\mathrm{sl}_3}^{*}\right) \rightarrow 0 \label{ldue}
\end{equation}
(where we used the obvious $A_3\ltimes T$-equivariant isomorphism $Z\simeq
A_3\times \mathbf{C}^{*}$); (\ref{luno}) shows that $\alpha ^2\in \ker
h^{\prime }$ and the reverse inclusion will be established if we show that
the push-forward $j_{*}$ is zero.
Consider the projectivization $\mathbf{P}\left( Diag_{\mathrm{sl}_3}\right)
\simeq \mathbf{P}^1$ of $Diag_{\mathrm{sl}_3}$. We have a cartesian diagram
\[
\begin{tabular}{lll}
$Z$ & $\stackrel{j}{\hookrightarrow }$ & $Diag_{\mathrm{sl}_3}\backslash
\left\{ 0\right\} $ \\
$^p\downarrow $ & & $\qquad \downarrow ^\pi $ \\
$Z^{\prime }$ & $\stackunder{j^{\prime }}{\hookrightarrow }$ & $\mathbf{P}%
\left( Diag_{\mathrm{sl}_3}\right) $%
\end{tabular}
\]
where
\[
Z^{\prime }=\left\{ \left[ 1,1\right] ,\left[ -2,1\right] ,\left[
1,-2\right] \right\} \simeq A_3
\]
$A_3\ltimes T$-equivariantly. Since
\[
j_{*}\circ p^{*}=\pi ^{*}\circ j_{*}^{\prime }
\]
and $p^{*}$ is obviously an isomorphism, it is enough to show that
\begin{equation}
im\left( j_{*}^{\prime }\right) \subseteq \ker \left( \pi ^{*}\right)
=\left( \ell \right) \subset \frac{A_{A_3\ltimes T}^{*}\left[ \ell \right] }{%
\left( \ell ^2-\alpha ^2\right) } \label{ltre}
\end{equation}
by the projective bundle theorem.
To compute $j_{*}^{\prime }$ we translate it into a transfer map. Consider
the $A_3\ltimes T$-equivariant commutative diagram
\[
\begin{tabular}{lll}
$Z^{\prime }\stackrel{\rho }{\longrightarrow }$ & $A_3\times \mathbf{P}%
\left( Diag_{\mathrm{sl}_3}\right) $ & \\
$^{j^{\prime }}\searrow $ & $\qquad \downarrow ^{pr_2}$ & \\
& $\mathbf{P}\left( Diag_{\mathrm{sl}_3}\right) $ &
\end{tabular}
\]
where ($\sigma =\left( 123\right) \in A_3$)
\begin{eqnarray*}
\rho \left( \left[ 1,1\right] \right) &=&\left( 1,\left[ 1,1\right] \right)
\\
\rho \left( \left[ -2,1\right] \right) &=&\left( \sigma ,\left[ -2,1\right]
\right) \\
\rho \left( \left[ 1,-2\right] \right) &=&\left( \sigma ^2,\left[
1,-2\right] \right) .
\end{eqnarray*}
Since
\[
A_{A_3\ltimes T}^{*}\left( A_3\times \mathbf{P}\left( Diag_{\mathrm{sl}%
_3}\right) \right) \simeq A_T^{*}\left( \mathbf{P}\left( Diag_{\mathrm{sl}%
_3}\right) \right) ,
\]
we have
\[
j_{*}^{\prime }\left( \xi \right) =pr_{2*}\circ \rho _{*}\left( \xi \right) =%
\mathrm{tsf}_T^{A_3\ltimes T}\left( \mathbf{P}\left( Diag_{\mathrm{sl}%
_3}\right) \right) \left( \xi \cdot \left[ \left\{ \left[ 1,1\right]
\right\} \right] \right)
\]
for any $\xi \in A_T^{*}\simeq A_{A_3\ltimes T}^{*}\left( Z^{\prime }\right)
$, where $\left\{ \left[ 1,1\right] \right\} $ is a $T$-invariant cycle on $%
\mathbf{P}\left( Diag_{\mathrm{sl}_3}\right) $.
Now, $\left\{ \left[ 1,1\right] \right\} $ is the zero scheme of the $T$%
-invariant regular section $\left( x_1-x_2\right) \in \Gamma \left( \mathbf{P%
}\left( Diag_{\mathrm{sl}_3}\right) ,\mathcal{O}\left( 1\right) \right) $,
therefore
\[
\left[ \left\{ \left[ 1,1\right] \right\} \right] =c_1\left( \mathcal{O}%
\left( 1\right) \right) \equiv \ell ^{\prime }\in A_T^{*}\left( \mathbf{P}%
\left( Diag_{\mathrm{sl}_3}\right) \right)
\]
and, obviously,
\[
\mathrm{res}_{A_3\ltimes T}^T\left( \mathbf{P}\left( Diag_{\mathrm{sl}%
_3}\right) \right) \left( \ell \right) =\ell ^{\prime }.
\]
By projection formula, we then get
\[
j_{*}^{\prime }\left( \xi \right) =\mathrm{tsf}_T^{A_3\ltimes T}\left(
\mathbf{P}\left( Diag_{\mathrm{sl}_3}\right) \right) \left( \xi \cdot \ell
^{\prime }\right) =\ell \cdot \mathrm{tsf}_T^{A_3\ltimes T}\left( \xi
\right)
\]
for any $\xi \in A_T^{*}\simeq A_{A_3\ltimes T}^{*}\left( Z^{\prime }\right)
$, which proves (\ref{ltre}).
(iii) By Prop. \ref{basico}, there are integers $A,$ $B$ such that
\[
c_2\left( W^\varepsilon \right) =A\alpha ^2+Bc_2\left( W\right) .
\]
Restricting this to $T$, we get $B=1$ and applying the involution $\left(
\cdot \right) ^\varepsilon $ we obtain $A\equiv 0$ $\limfunc{mod}3$.
Again by Prop. \ref{basico}, there are integers $A,B,C,D$ such that
\[
c_3\left( W^\varepsilon \right) =A\alpha ^3+B\alpha c_2\left( W\right)
+Cc_3\left( W\right) +D\theta
\]
in $A_{A_3\ltimes T}^{*}$. Restricting to $T$, we get
\[
\left( C+1\right) u_1u_2u_3+D\left( u_2^2u_3+u_3^2u_1+u_1^2u_2\right) =0\in
A_{T_{SL_3}}^{*}=\frac{\mathbf{Z}\left[ u_1,u_2,u_3\right] }{\left(
u_1+u_2+u_3\right) };
\]
but $u_2^2u_3+u_3^2u_1+u_1^2u_2$ and $u_1u_2u_3$ are linearly independent,
hence $C=-1,$ $D=0$. Now apply the involution $\left( \cdot \right)
^\varepsilon $ to get
\[
A\alpha ^3+B\alpha c_2\left( W\right) =0.
\]
Since (Remark \ref{twistaction})
\[
\theta ^\varepsilon =-\mathrm{tsf}_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}\left(
u_1^2u_3\right) ,
\]
an easy computation yields
\[
\left( \theta -\theta ^\varepsilon +3c_3\left( W\right) \right) _{\mid
T_{SL_3}}=0.
\]
Therefore (Proposition \ref{basico} and (i) of this Lemma) there exist $%
A,B,C\in \mathbf{Z}$ such that
\[
\theta -\theta ^\varepsilon +3c_3\left( W\right) =A\alpha ^3+Bc_3\left(
W\right) +C\theta
\]
is $3$-torsion. Then, restricting to $T_{SL_3}$ and observing that $%
c_3\left( W\right) _{\mid T}$ and $\theta _{\mid T}$ are linearly
independent, we get $B=C=0$; restricting now to $A_3$, we obtain $A\equiv 0$
$\limfunc{mod}3$ (since
\[
\mathrm{res}_{A_3\ltimes T_{SL_3}}^{A_3}\circ \mathrm{tsf}%
_{T_{SL_3}}^{A_3\ltimes T_{SL_3}}=0).
\]
The $C_2$-invariance of $\underline{\chi }$ is a consequence of the
transformation rules of $c_2\left( W\right) $, $c_3\left( W\right) $ and $%
\theta $.
(iv) Since $q$ is $3$-torsion, we may suppose $2$ inverted. We have $q_{\mid
T_{SL_3}}=0$ because $A_{T_{SL_3}}^{*}$ is torsion-free. It is not difficult
to verify that
\[
\left( 2\Theta _{\mid T_{SL_3}}+3c_3\left( W\right) _{\mid T_{SL_3}}\right)
^2+4c_2\left( W\right) _{\mid T_{SL_3}}^3+27c_3\left( W\right) _{\mid
T_{SL_3}}^2=0.
\]
Then it is enough to prove that the ideal $\mathcal{I}$ of relations between
\[
\left\{ c_2(W)_{\mid T_{SL_3}},c_3(W)_{\mid T_{SL_3}},\Theta _{\mid
T_{SL_3}}\right\}
\]
in $A_{T_{SL_3}}^{*}\left[ \frac 12\right] $ is generated by just this one.
Now, $\Theta _{\mid T_{SL_3}}=-\frac 32c_3(W)_{\mid T_{SL_3}}+\frac 12\delta
$, where $\delta =\left( u_1-u_2\right) \left( u_2-u_3\right) \left(
u_1-u_3\right) $, so we have to show that
\[
\mathcal{I}=\left( \delta ^2+4c_2(W)_{\mid T_{SL_3}}^3+27c_3(W)_{\mid
T_{SL_3}}^2\right) .
\]
Let $p\in Z\left[ \frac 12\right] \left[ X,Y,Z\right] $ with
\[
p\left( c_2(W)_{\mid T_{SL_3}},c_3(W)_{\mid T_{SL_3}},\delta \right) =0
\]
in $A_{T_{SL_3}}^{*}\left[ \frac 12\right] $. We have
\begin{equation}
p\left( X,Y,Z\right) =p_0\left( X,Y\right) +Zp_1\left( X,Y\right) \text{
\quad mod}\left( Z^2+4X^3+27Y^2\right) . \label{star}
\end{equation}
If we let $C_2=\left\{ 1,\varepsilon \right\} $ act on $A_{T_{Sl_3}}^{*}%
\left[ \frac 12\right] $ permuting $u_1$ and $u_2$, we get
\begin{eqnarray*}
\left( c_2(W)_{\mid T_{SL_3}}\right) ^\varepsilon &=&c_2(W)_{\mid T_{SL_3}}
\\
\left( c_3(W)_{\mid T_{SL_3}}\right) ^\varepsilon &=&c_3(W)_{\mid T_{SL_3}}
\\
\delta ^\varepsilon &=&-\delta
\end{eqnarray*}
and then
\[
p^\varepsilon =p\left( c_2(W)_{\mid T_{SL_3}},c_3(W)_{\mid T_{SL_3}},-\delta
\right) =0
\]
in $A_{T_{SL_3}}^{*}\left[ \frac 12\right] $ (note that $u_1+u_2+u_3$ is $C_2
$-invariant). From (\ref{star}) we get
\[
\left\{
\begin{array}{c}
p_0\left( c_2(W)_{\mid T_{SL_3}},c_3(W)_{\mid T_{SL_3}}\right) +p_1\left(
c_2(W)_{\mid T_{SL_3}},c_3(W)_{\mid T_{SL_3}}\right) \delta =0 \\
p_0\left( c_2(W)_{\mid T_{SL_3}},c_3(W)_{\mid T_{SL_3}}\right) -p_1\left(
c_2(W)_{\mid T_{SL_3}},c_3(W)_{\mid T_{SL_3}}\right) \delta =0
\end{array}
\right.
\]
so ($\delta \neq 0$)
\[
p_0\left( c_2(W)_{\mid T_{SL_3}},c_3(W)_{\mid T_{SL_3}}\right) =p_1\left(
c_2(W)_{\mid T_{SL_3}},c_3(W)_{\mid T_{SL_3}}\right) =0.
\]
But $c_2(W)_{\mid T_{SL_3}}$ and $c_3(W)_{\mid T_{SL_3}}$ are algebraically
independent, thus $p_0\left( X,Y\right) =p_1\left( X,Y\right) =0$ as
polynomials, as desired. \TeXButton{End Proof}{\endproof}
Therefore both $h^{\prime }\left( \alpha c_3(W)\right) $ and $h^{\prime
}\left( \underline{\chi }\right) $ can be identified (via lemma \ref
{transfertrick}) with their transfers, which are elements of $_3A_{\Gamma
_3}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right) $.
\begin{proposition}
\label{final}The natural morphism
\[
f:A_{\Gamma _3}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right) \longrightarrow
\left( A_T^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right) \right) ^{S_3}=\left(
A_T^{*}\right) ^{S_3}
\]
is surjective with kernel $\left( h^{\prime }\left( \alpha c_3\left(
W\right) \right) ,h^{\prime }\left( \underline{\chi }\right) \right) $,
where
\[
h^{\prime }:A_{A_3\ltimes T}^{*}\longrightarrow A_{A_3\ltimes T}^{*}\left(
Diag_{\mathrm{sl}_3}^{*}\right)
\]
is the pullback.
\end{proposition}
\TeXButton{Proof}{\proof} Commutativity of
\begin{tabular}{ll}
$A_{\Gamma _3}^{*}\stackrel{g}{\quad \longrightarrow }$ & $(A_T^{*}\left(
Diag_{\mathrm{sl}_3}^{*}\right) $ \\
$^h\downarrow $ & $\nearrow _f$ \\
$A_{\Gamma _3}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right) $ &
\end{tabular}
together with lemma \ref{hsurj}, prove that $f$ is surjective. Moreover $%
h^{\prime }\left( \alpha c_3\left( W\right) \right) $ and $h^{\prime }\left(
\underline{\chi }\right) $ are $3$-torsion so $\ker f\supseteq \left(
h^{\prime }\left( \alpha c_3\left( W\right) \right) ,h^{\prime }\left(
\underline{\chi }\right) \right) $ since $A_T^{*}$ is torsion-free. So we
are left to prove the reverse inclusion.
\texttt{CLAIM. }$\ker f=$ $_3A_{\Gamma _3}^{*}\left( Diag_{\mathrm{sl}%
_3}^{*}\right) \simeq $ $_3\left( A_{A_3\ltimes T}^{*}\left( Diag_{\mathrm{sl%
}_3}^{*}\right) \right) ^{C_2}$.
\underline{Proof of Claim}. $A_T^{*}$ is torsion-free, so $\ker f\supseteq $
$_3A_{\Gamma _3}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right) $. The pullback $%
\pi :A_{PGL_3}^{*}\rightarrow \left( A_T^{*}\right) ^{S_3}$ factors as
\[
A_{PGL_3}^{*}\stackrel{p}{\twoheadrightarrow }A_{PGL_3}^{*}\left( U\right)
\simeq A_{\Gamma _3}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right) \stackrel{f}{%
\longrightarrow }\left( A_T^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right)
\right) ^{S_3}=\left( A_T^{*}\right) ^{S_3}
\]
and from Prop. \ref{torsion}, we get $\ker \pi =$ $_3A_{PGL_3}^{*}$; so $%
\ker \left( f\circ p\right) =$ $_3A_{PGL_3}^{*}$ and we conclude since $p$
is surjective. $\mathbf{\Box }$
Now, let $\xi \in _3\left( A_{A_3\ltimes T}^{*}\left( Diag_{\mathrm{sl}%
_3}^{*}\right) \right) ^{C_2}$. Omitting to write $h^{\prime }\left( \cdot
\right) $ everywhere and denoting $\mathrm{tsf}_T^{A_3\ltimes T}\left(
u_2^2u_3\right) $ by $\theta $, we must have
\[
\xi =\alpha \cdot p\left( c_2(W),c_3(W)\right) +\underline{\chi }\cdot
q\left( c_2(W),c_3(W),\theta \right)
\]
for some polynomials $p$ and $q$, since $\xi $ is $3$-torsion (we used Prop.
\ref{basico} and lemma \ref{basico2} (i), (ii), (iv) ). But $\xi $ is also $%
C_2$-invariant, so if $C_2=\left\{ 1,\varepsilon \right\} ,$ we have:
\[
\xi =-2\xi =-\left( \xi +\xi ^\varepsilon \right) =\alpha \cdot \left(
p^\varepsilon -p\right) +\underline{\chi }\cdot \left( -\left(
q+q^\varepsilon \right) \right)
\]
(lemma \ref{basico2} (iii)). By lemma \ref{basico2} (iii), we have
\[
\alpha \cdot p^\varepsilon =\alpha \cdot p\left( c_2(W),-c_3(W)\right)
\]
and we can write $\alpha \left( p-p^\varepsilon \right) $ as a polynomial of
the form
\[
\alpha c_3\left( W\right) \cdot p^{\prime }\left( c_2(W),c_3(W)^2\right)
\]
for some polynomial $p^{\prime }$. Bt the \texttt{CLAIM} above, we conclude
that $\ker f\subseteq \left( \alpha c_3\left( W\right) ,\underline{\chi }%
\right) $. \TeXButton{End Proof}{\endproof}\
Let us summarize the situation so far. We are studying the first step (\ref
{tre}) of the stratification of $\mathrm{sl}_3$. So we started studying $%
A_{PGL_3}^{*}\left( U\right) $. We have an isomorphism $A_{PGL_3}^{*}\left(
U\right) \simeq A_{\Gamma _3}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right) $
(Prop. \ref{p3.1}) and an exact sequence (Prop. \ref{final}):
\[
0\rightarrow \left( \alpha c_3\left( W\right) ,\underline{\chi }\right)
\longrightarrow A_{\Gamma _3}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right)
\stackrel{f}{\longrightarrow }\left( A_T^{*}\left( Diag_{\mathrm{sl}%
_3}^{*}\right) \right) ^{S_3}=\left( A_T^{*}\right) ^{S_3}\rightarrow 0.
\]
To be precise, $\alpha c_3\left( W\right) $ and $\underline{\chi }\ $belong
to $A_{A_3\ltimes T}^{*}$ but we denote by the same symbols the elements
\begin{eqnarray*}
&&\ \ \ \left( \mathrm{tsf}_{A_3\ltimes T}^{\Gamma _3}\left( Diag_{\mathrm{sl%
}_3}^{*}\right) \circ h^{\prime }\right) \left( \alpha c_3\left( W\right)
\right) \\
&&\ \ \ \left( \mathrm{tsf}_{A_3\ltimes T}^{\Gamma _3}\left( Diag_{\mathrm{sl%
}_3}^{*}\right) \circ h^{\prime }\right) \left( \underline{\chi }\right)
\end{eqnarray*}
in $A_{\Gamma _3}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right) $, where
\[
\mathrm{tsf}_{A_3\ltimes T}^{\Gamma _3}\left( Diag_{\mathrm{sl}%
_3}^{*}\right) :A_{A_3\ltimes T}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right)
\rightarrow A_{\Gamma _3}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right)
\]
is the transfer morphism and
\[
h^{\prime }:A_{A_3\ltimes T}^{*}\longrightarrow A_{A_3\ltimes T}^{*}\left(
Diag_{\mathrm{sl}_3}^{*}\right)
\]
is the obvious pullback. Moreover, by the proof of lemma \ref{hsurj} and
with the same notations, the elements
\[
\left\{ 2c_2\left( \mathrm{sl}_3\right) -c_2\left( Sym^3E\right) ,c_3\left(
Sym^3E\right) ,c_6\left( \mathrm{sl}_3\right) \right\} \subset A_{PGL_3}^{*}
\]
project to the three generators (lemma \ref{invariants}) of $\left(
A_T^{*}\right) ^{S_3}$ through the composition
\[
A_{PGL_3}^{*}\rightarrow A_{PGL_3}^{*}\left( U\right) \simeq A_{\Gamma
_3}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right) \stackrel{f}{\longrightarrow }%
\left( A_T^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right) \right) ^{S_3}=\left(
A_T^{*}\right) ^{S_3}\text{.}
\]
If we lift the elements
\[
\ \ \ \left( \mathrm{tsf}_{A_3\ltimes T}^{\Gamma _3}\left( Diag_{\mathrm{sl}%
_3}^{*}\right) \circ h^{\prime }\right) \left( \alpha c_3\left( W\right)
\right) ,\text{ }
\]
\[
\ \left( \mathrm{tsf}_{A_3\ltimes T}^{\Gamma _3}\left( Diag_{\mathrm{sl}%
_3}^{*}\right) \circ h^{\prime }\right) \left( \underline{\chi }\right) \in
A_{\Gamma _3}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right) ,
\]
respectively to elements $\rho ,\chi \in A_{PGL_3}^{*}$, via the surjective
pullback
\[
A_{PGL_3}^{*}\longrightarrow A_{PGL_3}^{*}\left( U\right) \simeq A_{\Gamma
_3}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right) \text{,}
\]
we find the following $5$ generators of $A_{PGL_3}^{*}$ coming from the open
subscheme $U\subset \mathrm{sl}_3$ (through the first step (\ref{tre}) of
the stratification of $\mathrm{sl}_3$)
\begin{equation}
\left\{ 2c_2\left( \mathrm{sl}_3\right) -c_2\left( Sym^3E\right) ,c_3\left(
Sym^3E\right) ,\rho ,\chi ,c_6\left( \mathrm{sl}_3\right) \right\} \text{,}
\label{generatorsfromU}
\end{equation}
with $\deg \rho =4$ and $\deg \chi =6$.
In the following subsection we will determine the other generators of $%
A_{PGL_3}^{*}$ coming from the complement $\mathrm{sl}_3\backslash U$,
starting from $Z_{1,1}$.
\subsection{Generators coming from the complement of $U\subset \mathrm{sl}_3$%
}
Consider again the first step of the stratification (\ref{due}):
\begin{equation}
A_{PGL_3}^{*}\left( Z_{1,1}\right) \stackrel{\left( j_{1,1}\right) _{*}}{%
\longrightarrow }A_{PGL_3}^{*}\left( \mathrm{sl}_3\backslash (Z_{1,0}\cup
Z_0\cup \left\{ 0\right\} )\right) \stackrel{i_{1,1}^{*}}{\longrightarrow }%
A_{PGL_3}^{*}\left( U\right) \rightarrow 0 \label{tre}
\end{equation}
where $\left( j_{1,1}\right) _{*}$ has degree $1$, equal to the codimension
of $Z_{1,1}$ in $\mathrm{sl}_3$.
\begin{lemma}
If $A\in Z_{1,1}$, let $g\in PGL_3$ be such that
\[
g^{-1}Ag=\left(
\begin{array}{ccc}
\lambda & 0 & 0 \\
1 & \lambda & 0 \\
0 & 0 & -2\lambda
\end{array}
\right) ;
\]
then, the rule
\[
A\longmapsto \left( \lambda ,\left[ g\right] \right)
\]
defines a $PGL_3$-equivariant isomorphism $Z_{1,1}\rightarrow \mathbf{A}%
^1\backslash \left\{ 0\right\} \times \frac{PGL_3}{\mathrm{U}_2\times
\mathbf{G}_m}$, where $\mathrm{U}_2$ is the full unipotent subgroup of $GL_2$
and $PGL_3$ acts trivially on $\mathbf{A}^1\backslash \left\{ 0\right\} $.
\end{lemma}
\TeXButton{Proof}{\proof} Everything is a straightforward verification left
to the interested reader. We only note that the stabilizer of
\[
\left(
\begin{array}{ccc}
\lambda & 0 & 0 \\
1 & \lambda & 0 \\
0 & 0 & -2\lambda
\end{array}
\right)
\]
(under the adjoint action of $PGL_3$) is
\[
\left\{ \left[ g\right] \mid g=\left(
\begin{array}{ccc}
\alpha & 0 & 0 \\
\beta & \alpha & 0 \\
0 & 0 & \gamma
\end{array}
\right) \text{, }\alpha ,\gamma \in \mathbf{G}_m\right\}
\]
which is obviously isomorphic to $\mathrm{U}_2\times \mathbf{G}_m$.
\TeXButton{End Proof}{\endproof}
By Corollary \ref{corunip}, Prop. \ref{productbygm}, \ref{homogeneous} and
lemma \ref{utrivial}, we have
\begin{equation}
A_{PGL_3}^{*}\left( Z_{1,1}\right) \simeq A_{\mathbf{G}_m}^{*}=\mathbf{Z}%
\left[ u\right] . \label{z11}
\end{equation}
It is not difficult to verify that
\[
j_{1,1}^{*}\left( 2c_2\left( \mathrm{sl}_3\right) -c_2\left( Sym^3E\right)
\right) =u^2\text{ ,}
\]
where we abused notation writing $2c_2\left( \mathrm{sl}_3\right) -c_2\left(
Sym^3E\right) $ for its pullback to $A_{PGL_3}^{*}\left( \mathrm{sl}%
_3\backslash (Z_{1,0}\cup Z_0\cup \left\{ 0\right\} )\right) $. Moreover
\[
\left( j_{1,1}\right) _{*}\left( 1\right) =\left[ Z_{1,1}\right]
=D^{*}\left( \left[ \left\{ 0\right\} \right] \right) =0
\]
where $D:\mathrm{sl}_3\backslash (Z_{1,0}\cup Z_0\cup \left\{ 0\right\}
)\rightarrow \mathbf{A}^1$ is the discriminant; so, by projection formula,
the ideal $\mathrm{im}\left( j_{1,1}\right) _{*}$ is generated by $\left(
j_{1,1}\right) _{*}\left( u\right) $.
Let $\Theta _{1,1}^{\left( 2\right) }$ be a lift of $\left( j_{1,1}\right)
_{*}\left( u\right) \in A_{PGL_3}^{*}\left( \mathrm{sl}_3\backslash
(Z_{1,0}\cup Z_0\cup \left\{ 0\right\} )\right) $ to $A_{PGL_3}^{*}$. The
analysis we made of (\ref{tre}) has the following upshot (recall (\ref
{generatorsfromU})): $A_{PGL_3}^{*}\left( \mathrm{sl}_3\backslash
(Z_{1,0}\cup Z_0\cup \left\{ 0\right\} )\right) $\emph{\ is generated by
(the images via }$A_{PGL_3}^{*}\twoheadrightarrow A_{PGL_3}^{*}\left(
\mathrm{sl}_3\backslash (Z_{1,0}\cup Z_0\cup \left\{ 0\right\} )\right) $%
\emph{\ of) }
\begin{equation}
\left\{ 2c_2\left( \mathrm{sl}_3\right) -c_2\left( Sym^3E\right) ,\Theta
_{1,1}^{\left( 2\right) },c_3\left( Sym^3E\right) ,\rho ,\chi ,c_6\left(
\mathrm{sl}_3\right) \right\} \text{.} \label{firststep}
\end{equation}
\
Now let us proceed one step further in the analysis of stratification (\ref
{due}); the second exact sequence of (\ref{uno}) is:
\begin{equation}
A_{PGL_3}^{*}\left( Z_{1,0}\right) \stackrel{\left( j_{1,0}\right) _{*}}{%
\longrightarrow }A_{PGL_3}^{*}\left( \mathrm{sl}_3\backslash (Z_0\cup
\left\{ 0\right\} )\right) \stackrel{i_{1,0}^{*}}{\longrightarrow }%
A_{PGL_3}^{*}\left( \mathrm{sl}_3\backslash (Z_{1,0}\cup Z_0\cup \left\{
0\right\} )\right) \rightarrow 0 \label{quattro}
\end{equation}
where $\left( j_{1,0}\right) _{*}$ has degree $3$, equal to the codimension
of $Z_{1,0}$ in $\mathrm{sl}_3$. We omit the straightforward proof of the
following:
\begin{lemma}
If $A\in Z_{1,0}$, let $g\in PGL_3$ be such that
\[
g^{-1}Ag=\left(
\begin{array}{ccc}
\lambda & 0 & 0 \\
0 & \lambda & 0 \\
0 & 0 & -2\lambda
\end{array}
\right) .
\]
Then, the rule
\[
A\longmapsto \left( \lambda ,\left[ g\right] \right)
\]
defines a $PGL_3$-equivariant isomorphism $Z_{1,0}\rightarrow \mathbf{A}%
^1\backslash \left\{ 0\right\} \times \frac{PGL_3}{GL_2}$, where $GL_2$
injects as
\[
\left(
\begin{array}{cc}
GL_2 & 0 \\
0 & 1
\end{array}
\right)
\]
and $PGL_3$ acts trivially on $\mathbf{A}^1\backslash \left\{ 0\right\} $.
\end{lemma}
Then, by Prop. \ref{homogeneous} and lemma \ref{utrivial}, we have\footnote{$%
\lambda _i\doteq c_i\left( \text{\textrm{standard representation}}\right) $.}
\begin{equation}
A_{PGL_3}^{*}\left( Z_{1,0}\right) \simeq A_{GL_2}^{*}=\mathbf{Z}\left[
\lambda _1,\lambda _2\right] \text{.} \label{az10}
\end{equation}
\begin{lemma}
\label{diga}$\left( j_{1,0}\right) _{*}$ is $3$-torsion.
\end{lemma}
\TeXButton{Proof}{\proof}If $\xi \in A_{PGL_3}^{*}\left( Z_{1,0}\right) $,
let $\widehat{\xi }\in A_{PGL_3}^{*}$ be a lift of $\left( j_{1,0}\right)
_{*}\left( \xi \right) $ via the surjective pullback
\[
\pi _{1,0}:A_{PGL_3}^{*}\longrightarrow A_{PGL_3}^{*}\left( sl_3\backslash
Z_0\cup \left\{ 0\right\} \right) .
\]
It is enough to prove that $\widehat{\xi }$ is $3$-torsion i.e. that
\[
\widehat{\xi }\in \ker \left( A_{PGL_3}^{*}\longrightarrow \left(
A_T^{*}\right) ^{S_3}\right) ,
\]
since by \cite{EG1}, Prop. 6, the rational pullback
\[
A_{PGL_3}^{*}\otimes \mathbf{Q}\longrightarrow \left( A_T^{*}\right)
^{S_3}\otimes \mathbf{Q}
\]
is an isomorphism and $A_{PGL_3}^{*}$ has only $3$-torsion by Cor. \ref
{cortorsion}.
Now, observe that
\[
\left( j_{1,0}\right) _{*}\left( \xi \right) \in \ker \left(
A_{PGL_3}^{*}\left( \mathrm{sl}_3\backslash Z_0\cup \left\{ 0\right\}
\right) \rightarrow A_{PGL_3}^{*}\left( \mathrm{sl}_3\backslash Z_{1,0}\cup
Z_0\cup \left\{ 0\right\} \right) \right)
\]
by the obvious localization sequence and therefore
\[
\widehat{\xi }\in \ker \left( A_{PGL_3}^{*}\longrightarrow
A_{PGL_3}^{*}\left( U\right) \right) ,
\]
by (\ref{tre}). To conclude, we note that $A_{PGL_3}^{*}\longrightarrow
\left( A_T^{*}\right) ^{S_3}$ factors as
\[
A_{PGL_3}^{*}\longrightarrow A_{PGL_3}^{*}\left( U\right) \simeq A_{\Gamma
_3}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right) \longrightarrow \left(
A_T^{*}\right) ^{S_3}\text{.}
\]
\TeXButton{End Proof}{\endproof}
\begin{proposition}
The ideal $\mathrm{im}\left( j_{1,0}\right) _{*}$ is generated by
\[
\left\{ \left( j_{1,0}\right) _{*}\left( 1\right) ,\left( j_{1,0}\right)
_{*}\left( \lambda _1\right) ,\left( j_{1,0}\right) _{*}\left( \lambda
_2\right) ,\left( j_{1,0}\right) _{*}\left( \lambda _2^2\right) ,\left(
j_{1,0}\right) _{*}\left( \lambda _1\lambda _2\right) ,\left( j_{1,0}\right)
_{*}\left( \lambda _1\lambda _2^2\right) \right\} .
\]
\end{proposition}
\TeXButton{Proof}{\proof}Identifying $A_{PGL_3}^{*}\left( Z_{1,0}\right) \ $%
with $A_{GL_2}^{*}=\mathbf{Z}\left[ \lambda _1,\lambda _2\right] $ via (\ref
{az10}) and writing $\left( \cdot \right) _{\mid GL_2}$ for $j_{1,0}^{*}$,
one can easily verify that
\begin{equation}
\left( \lambda \equiv 2c_2\left( \mathrm{sl}_3\right) -c_2\left(
Sym^3E\right) \right) _{\mid GL_2}=\lambda _1^2-3\lambda _2\doteq \tau _2
\label{alef}
\end{equation}
\begin{equation}
c_6\left( \mathrm{sl}_3\right) _{\mid GL_2}=-\lambda _1^2\lambda
_2^2+4\lambda _2^3\doteq \tau _6. \label{alef1}
\end{equation}
Therefore
\begin{equation}
\lambda _2^3=\tau _6+\tau _2\lambda _2^2, \label{alef3}
\end{equation}
and if $\frak{I}$ denotes the ideal generated by
\[
\begin{array}{c}
\left\{ \left( j_{1,0}\right) _{*}\left( 1\right) ,\left( j_{1,0}\right)
_{*}\left( \lambda _1\right) ,\left( j_{1,0}\right) _{*}\left( \lambda
_2\right) ,\right. \\
\left. \left( j_{1,0}\right) _{*}\left( \lambda _2^2\right) ,\left(
j_{1,0}\right) _{*}\left( \lambda _1\lambda _2\right) ,\left( j_{1,0}\right)
_{*}\left( \lambda _1\lambda _2^2\right) \right\} ,
\end{array}
\]
we have
\begin{equation}
\left( j_{1,0}\right) _{*}\left( \lambda _2^m\right) \in \frak{I},\text{ }%
\forall m\geq 0, \label{alef2}
\end{equation}
by an easy induction on $m$, using projection formula.
Now, consider the general monomial $\lambda _1^n\lambda _2^m$. If $n=2r$, we
have
\begin{equation}
\lambda _1^n\lambda _2^m=\left( \tau _2+3\lambda _2\right) ^r\lambda
_2^m\equiv \tau _2^r\lambda _2^m\text{ }\left( \limfunc{mod}3\right)
\label{alef4}
\end{equation}
and then
\[
\left( j_{1,0}\right) _{*}\left( \lambda _1^{2r}\lambda _2^m\right) =\lambda
^r\cdot \left( j_{1,0}\right) _{*}\left( \lambda _2^m\right) ,
\]
by lemma \ref{diga} and projection formula; thus
\[
\left( j_{1,0}\right) _{*}\left( \lambda _1^{2r}\lambda _2^m\right) \in
\frak{I},\text{ }\forall m,r\geq 0,
\]
by (\ref{alef2}). If $n=2r+1$, (\ref{alef4}), (\ref{alef3}) and projection
formula easily reduce the assert
\[
\left( j_{1,0}\right) _{*}\left( \lambda _1^{2r+1}\lambda _2^m\right) \in
\frak{I},\text{ }\forall m,r\geq 0
\]
to the assert
\[
\left( j_{1,0}\right) _{*}\left( \lambda _1\lambda _2^m\right) \in \frak{I},%
\text{ }\forall m\geq 0,
\]
which is easily proved by induction on $m$.
Since the monomials $\lambda _1^n\lambda _2^m$ generates $A_{GL_2}^{*}$ as a
$\mathbf{Z}$-module, we conclude that
\[
\frak{I}=\mathrm{im}\left( j_{1,0}\right) _{*}\text{ .}
\]
\TeXButton{End Proof}{\endproof}
Therefore, if we denote by $\Theta _{1,0}^{\left( 3\right) }$ (respectively,
$\Theta _{1,0}^{\left( 4\right) }$, $\Theta _{1,0}^{\left( 5\right) }$, $%
\Theta _{1,0}^{\left( 6\right) }$, $\Theta _{1,0}^{\left( 7\right) }$, $%
\Theta _{1,0}^{\left( 8\right) }$) a lift of $\left( j_{1,0}\right)
_{*}\left( 1\right) $ (respectively, of $\left( j_{1,0}\right) _{*}\left(
\lambda _1\right) $, $\left( j_{1,0}\right) _{*}\left( \lambda _2\right) $, $%
\left( j_{1,0}\right) _{*}\left( \lambda _1\lambda _2\right) $, $\left(
j_{1,0}\right) _{*}\left( \lambda _2^2\right) $, $\left( j_{1,0}\right)
_{*}\left( \lambda _1\lambda _2^2\right) $) to $A_{PGL_3}^{*}$, from (\ref
{firststep}) and (\ref{quattro}) we get that $A_{PGL_3}^{*}\left( \mathrm{sl}%
_3\backslash (Z_0\cup \left\{ 0\right\} )\right) $\emph{\ is generated by
(the images via }$A_{PGL_3}^{*}\twoheadrightarrow A_{PGL_3}^{*}\left(
\mathrm{sl}_3\backslash (Z_0\cup \left\{ 0\right\} )\right) $\emph{\ of) }
\begin{equation}
\begin{array}{c}
\left\{ 2c_2\left( \mathrm{sl}_3\right) -c_2\left( Sym^3E\right) ,\Theta
_{1,1}^{\left( 2\right) },c_3\left( Sym^3E\right) ,\Theta _{1,0}^{\left(
3\right) },\right. \\
\left. \rho ,\Theta _{1,0}^{\left( 4\right) },\Theta _{1,0}^{\left( 5\right)
},\chi ,\Theta _{1,0}^{\left( 6\right) },c_6\left( \mathrm{sl}_3\right)
,\Theta _{1,0}^{\left( 7\right) },\Theta _{1,0}^{\left( 8\right) }\right\} .
\end{array}
\label{secondstep}
\end{equation}
Let us proceed one step further in the analysis of stratification (\ref{due}%
); the third exact sequence of (\ref{uno}), in our case is:
\begin{equation}
A_{PGL_3}^{*}\left( Z_{0,1}\right) \stackrel{\left( j_{0,1}\right) _{*}}{%
\longrightarrow }A_{PGL_3}^{*}\left( \mathrm{sl}_3\backslash (Z_{0,0}\cup
\left\{ 0\right\} )\right) \stackrel{i_{0,1}^{*}}{\longrightarrow }%
A_{PGL_3}^{*}\left( \mathrm{sl}_3\backslash (Z_0\cup \left\{ 0\right\}
)\right) \rightarrow 0 \label{cinque}
\end{equation}
where $\left( j_{0,1}\right) _{*}$ has degree $2$, equal to the codimension
of $Z_{0,1}$ in $\mathrm{sl}_3$. $Z_{0,1}$ is a $PGL_3$-orbit with
stabilizer
\[
\left\{ \left[ g\right] \mid g=\left(
\begin{array}{ccc}
1 & 0 & 0 \\
\alpha & 1 & 0 \\
\beta & \alpha & 1
\end{array}
\right) ,\alpha ,\beta \in \mathbf{C}\right\}
\]
which is unipotent and then, by Prop. \ref{unipotent}, we have $%
A_{PGL_3}^{*}\left( Z_{0,1}\right) =\mathbf{Z}$. If we denote by $\Theta
_{0,1}^{\left( 2\right) }$ a lift of $\left( j_{0,1}\right) _{*}\left(
1\right) =\left[ Z_{0,1}\right] \in A_{PGL_3}^2\left( \mathrm{sl}%
_3\backslash (Z_{0,0}\cup \left\{ 0\right\} )\right) $ via the surjective
pullback
\[
A_{PGL_3}^{*}\longrightarrow A_{PGL_3}^{*}\left( \mathrm{sl}_3\backslash
(Z_{0,0}\cup \left\{ 0\right\} )\right) ,
\]
from (\ref{cinque}) and (\ref{secondstep}) we get that $A_{PGL_3}^{*}\left(
\mathrm{sl}_3\backslash (Z_{0,0}\cup \left\{ 0\right\} )\right) $\emph{\ is
generated by (the images via }$A_{PGL_3}^{*}\twoheadrightarrow
A_{PGL_3}^{*}\left( \mathrm{sl}_3\backslash (Z_{0,0}\cup \left\{ 0\right\}
)\right) $\emph{\ of) }
\begin{equation}
\begin{array}{c}
\left\{ 2c_2\left( \mathrm{sl}_3\right) -c_2\left( Sym^3E\right) ,\Theta
_{1,1}^{\left( 2\right) },\Theta _{0,1}^{\left( 2\right) },c_3\left(
Sym^3E\right) ,\Theta _{1,0}^{\left( 3\right) },\right. \\
\left. \rho ,\Theta _{1,0}^{\left( 4\right) },\Theta _{1,0}^{\left( 5\right)
},\chi ,\Theta _{1,0}^{\left( 6\right) },c_6\left( \mathrm{sl}_3\right)
,\Theta _{1,0}^{\left( 7\right) },\Theta _{1,0}^{\left( 8\right) }\right\} .
\end{array}
\label{secondbisstep}
\end{equation}
We have come to the second-last step of stratification (\ref{due}):
\begin{equation}
A_{PGL_3}^{*}\left( Z_{0,0}\right) \stackrel{\left( j_{0,0}\right) _{*}}{%
\longrightarrow }A_{PGL_3}^{*}\left( \mathrm{sl}_3\backslash \left\{
0\right\} \right) \stackrel{i_{0,0}^{*}}{\longrightarrow }%
A_{PGL_3}^{*}\left( \mathrm{sl}_3\backslash (Z_{0,0}\cup \left\{ 0\right\}
)\right) \rightarrow 0 \label{sei}
\end{equation}
where $\left( j_{0,0}\right) _{*}$ has degree $4$, equal to the codimension
of $Z_{0,0}$ in $\mathrm{sl}_3$. $Z_{0,0}$ is a $PGL_3$-orbit with
stabilizer
\[
\left\{ \left[ g\right] \mid g=\left(
\begin{array}{ccc}
1 & 0 & 0 \\
\alpha & \delta & 0 \\
\beta & \gamma & 1
\end{array}
\right) ,\alpha ,\beta ,\gamma \in \mathbf{C,}\text{ }\delta \in \mathbf{G}%
_m\right\}
\]
which is a split extension of $\mathbf{G}_m$ by the full unipotent group $%
\mathrm{U}_3\subset GL_3$. By Cor. \ref{corunip} we get an isomorphism $%
A_{PGL_3}^{*}\left( Z_{0,0}\right) \simeq A_{\mathbf{G}_m}^{*}=\mathbf{Z}%
\left[ u\right] $. Since
\[
j_{0,0}^{*}\left( 2c_2\left( sl_3\right) -c_2\left( Sym^3E\right) \right)
=u^2,
\]
$A_{PGL_3}^{*}\left( Z_{0,0}\right) $ is generated by $\left\{ \left(
j_{0,0}\right) _{*}\left( 1\right) ,\left( j_{0,0}\right) _{*}\left(
u\right) \right\} $ as an $A_{PGL_3}^{*}\left( \mathrm{sl}_3\backslash
\left\{ 0\right\} \right) $-module (by projection formula) and if we denote
by $\Theta _{0,0}^{\left( 4\right) }$ (respectively, $\Theta _{0,0}^{\left(
5\right) }$) a lift of $\left( j_{0,0}\right) _{*}\left( 1\right) $
(respectively, of $\left( j_{0,0}\right) _{*}\left( u\right) $) to $%
A_{PGL_3}^{*}$, we get that $A_{PGL_3}^{*}\left( \mathrm{sl}_3\backslash
\left\{ 0\right\} \right) $\emph{\ is generated by (the images via }$%
A_{PGL_3}^{*}\twoheadrightarrow A_{PGL_3}^{*}\left( \mathrm{sl}_3\backslash
\left\{ 0\right\} \right) $\emph{\ of) }
\begin{equation}
\begin{array}{c}
\left\{ 2c_2\left( \mathrm{sl}_3\right) -c_2\left( Sym^3E\right) ,\Theta
_{1,1}^{\left( 2\right) },\Theta _{0,1}^{\left( 2\right) },c_3\left(
Sym^3E\right) ,\Theta _{1,0}^{\left( 3\right) },\right. \\
\left. \rho ,\Theta _{1,0}^{\left( 4\right) },\Theta _{0,0}^{(4)},\Theta
_{1,0}^{\left( 5\right) },\Theta _{0,0}^{(5)},\chi ,\Theta _{1,0}^{\left(
6\right) },c_6\left( \mathrm{sl}_3\right) ,\Theta _{1,0}^{\left( 7\right)
},\Theta _{1,0}^{\left( 8\right) }\right\} .
\end{array}
\label{thirdstep}
\end{equation}
The last step of (\ref{uno}) for stratification (\ref{due}) is immediate
because
\[
A_{PGL_3}^{*}\left( \mathrm{sl}_3\backslash \left\{ 0\right\} \right) \simeq
A_{PGL_3}^{*}\diagup \left( c_8\left( \mathrm{sl}_3\right) \right)
\]
by self-intersection formula (\cite{Fu}, p. 103).
Therefore we conclude our analysis of the stratification (\ref{due}) with
the following result:
\begin{proposition}
\label{prelgen}$A_{PGL_3}^{*}$\ is generated by
\begin{equation}
\begin{array}{c}
\left\{ 2c_2\left( \mathrm{sl}_3\right) -c_2\left( Sym^3E\right) ,\Theta
_{1,1}^{\left( 2\right) },\Theta _{0,1}^{\left( 2\right) },c_3\left(
Sym^3E\right) ,\Theta _{1,0}^{\left( 3\right) },\right. \\
\left. \rho ,\Theta _{1,0}^{\left( 4\right) },\Theta _{0,0}^{(4)},\Theta
_{1,0}^{\left( 5\right) },\Theta _{0,0}^{(5)},\chi ,\Theta _{1,0}^{\left(
6\right) },c_6\left( \mathrm{sl}_3\right) ,\Theta _{1,0}^{\left( 7\right)
},\Theta _{1,0}^{\left( 8\right) },c_8\left( \mathrm{sl}_3\right) \right\} ,
\end{array}
\label{laststep}
\end{equation}
where $\deg \Theta _{0,1}^{\left( 2\right) }=2,$ $\deg \Theta _{1,1}^{\left(
2\right) }=2$, $\deg \rho =4$, $\deg \chi =6$, $\deg \Theta _{1,0}^{\left(
m\right) }=m$\ and $\deg \Theta _{0,0}^{(r)}=r.$
\end{proposition}
We will make this result more precise in the following section by getting
rid of all the $\Theta $ generators.
\section{$A_{PGL_3}^{*}$ is not generated by Chern classes. Elimination of
some generators}
In this section we first prove that $A_{PGL_3}^{*}$ is not generated by
Chern classes and then that all its $\Theta $ generators are zero.
\begin{lemma}
\label{cohom}Writing $H_{PGL_3}^i$ for $H^i\left( \mathrm{B}PGL_3,\mathbf{Z}%
\right) $, we have:
\begin{equation}
\begin{tabular}{|ll|}
\hline
\multicolumn{1}{|l|}{$H_{PGL_3}^0\simeq \mathbf{Z}$} & $H_{PGL_3}^1=0$ \\
\hline
\multicolumn{1}{|l|}{$H_{PGL_3}^2=0$} & $H_{PGL_3}^3\simeq \mathbf{Z}/3$ \\
\hline
\multicolumn{1}{|l|}{$H_{PGL_3}^4\simeq \mathbf{Z}$} & $H_{PGL_3}^5=0$ \\
\hline
\multicolumn{1}{|l|}{$H_{PGL_3}^6\simeq \mathbf{Z}$} & $H_{PGL_3}^7=0$ \\
\hline
\multicolumn{1}{|l|}{$H_{PGL_3}^8\simeq \mathbf{Z\oplus Z}/3$} & $%
H_{PGL_3}^9=0$ \\ \hline
\multicolumn{1}{|l|}{$H_{PGL_3}^{10}\simeq \mathbf{Z}$} & $%
H_{PGL_3}^{11}\simeq \mathbf{Z}/3$ \\ \hline
\multicolumn{1}{|l|}{$H_{PGL_3}^{12}\simeq \mathbf{Z\oplus Z}$} & $%
H_{PGL_3}^{13}=0$ \\ \hline
\multicolumn{1}{|l|}{$H_{PGL_3}^{14}\simeq \mathbf{Z}$} & $%
H_{PGL_3}^{15}\simeq \mathbf{Z}/3$ \\ \hline
\multicolumn{1}{|l|}{$H_{PGL_3}^{16}\simeq \mathbf{Z\oplus Z}\oplus \mathbf{Z%
}/3$} & $\cdots $ \\ \hline
\end{tabular}
\label{Hstar}
\end{equation}
\end{lemma}
\TeXButton{Proof}{\proof} It is a routine computation using the Universal
Coefficients' Formula for cohomology (e.g. \cite{Sp}), once one knows the
following facts:
\begin{enumerate}
\item[1.] $H^{*}\left( \mathrm{B}PGL_3,\mathbf{Q}\right) \simeq H^{*}\left(
\mathrm{B}SL_3,\mathbf{Q}\right) =\mathbf{Q}\left[ c_2\left( E\right)
,c_3\left( E\right) \right] $, $E$ being the standard representation of $SL_3
$;
\item[2.] $H^{*}\left( \mathrm{B}PGL_3,\mathbf{Z}\right) $ has only $3$%
-torsion;
\item[3.] there is a ring isomorphism
\[
H^{*}\left( \mathrm{B}PGL_3,\mathbf{Z}/3\right) \simeq \mathbf{Z}/3\left[
y_2,y_8,y_{12}\right] \otimes \Lambda \left( y_3,y_7\right) \diagup \left(
y_2y_3,y_2y_7,y_2y_8+y_3y_7\right)
\]
where $\deg y_i=i$.
\end{enumerate}
1. follows immediately from the Leray spectral sequence
\[
H^p\left( \mathrm{B}PGL_3,H^q\left( \mathrm{B}\mathbf{\mu }_3,\mathbf{Z}%
\right) \right) \Longrightarrow H^{p+q}\left( \mathrm{B}SL_3,\mathbf{Z}%
\right) ;
\]
2. is proved in \cite{KY}, p. 790 and 3. was computed in \cite{KMS}.
\TeXButton{End Proof}{\endproof}
\begin{theorem}
\label{result1}$A_{PGL_3}^{*}$ is not generated by Chern classes; more
precisely, $\rho $ is not a polynomial in Chern classes.
\end{theorem}
\TeXButton{Proof}{\proof} We proceed in 4 steps:
\begin{enumerate}
\item[(I)] First we show that $\mathrm{cl}\left( \rho \right) $ is nonzero
in $H^8(\mathrm{B}PGL_3,\mathbf{Z})_{tors}$, where $\mathrm{cl}%
:A_{PGL_3}^{*}\rightarrow H^{*}(\mathrm{B}PGL_3,\mathbf{Z})$ is the cycle
class map and $\rho $ is one of the generators of $A_{PGL_3}^{*}$ (see Prop.
\ref{prelgen});
\item[(II)] Then we use a spectral sequence argument to show that
\[
\mathrm{im}(H^8(\mathrm{B}PGL_3,\mathbf{Z})\rightarrow H^8(\mathrm{B}SL_3,%
\mathbf{Z}))
\]
has index at least $9$ in $H^8(\mathrm{B}SL_3,\mathbf{Z})\simeq \mathbf{Z}$;
\item[(III)] Next, we use the fact that $c_2\left( \mathrm{sl}_3\right)
^2\mapsto 36\alpha _2^2$ via
\[
H^8(\mathrm{B}PGL_3,\mathbf{Z})\simeq \mathbf{Z}\oplus \mathbf{Z}%
/3\longrightarrow H^8(\mathrm{B}SL_3,\mathbf{Z})\simeq \mathbf{Z\cdot }%
\alpha _2^2
\]
to conclude that
\[
H^8(\mathrm{B}PGL_3,\mathbf{Z}_{(3)})\simeq \mathbf{Z}_{(3)}\cdot c_2\left(
\mathrm{sl}_3\right) ^2\oplus \mathbf{Z}/3\cdot \mathrm{cl}(\rho )
\]
(where we have written $\alpha $ in place of $j_{\bullet }\left( \alpha
\right) $, with $j_{\bullet }:H^{*}(\mathrm{B}PGL_3,\mathbf{Z})\rightarrow
H^{*}(\mathrm{B}PGL_3,\mathbf{Z}_{(3)})$ induced by the localization $j:%
\mathbf{Z}\rightarrow \mathbf{Z}_{(3)}$ and $\alpha _i\doteq c_i\left( \text{%
standard repr. of }SL_3\right) $);
\item[(IV)] Finally we show that $\mathrm{cl}(\rho )\in H^8(\mathrm{B}PGL_3,%
\mathbf{Z})$ is not in the Chern subring of $H^{*}(\mathrm{B}PGL_3,\mathbf{Z}%
)$ (implying that $\rho $ itself is not in the Chern subring of $%
A_{PGL_3}^{*}$).
\end{enumerate}
(I) We freely use Remark \ref{twistaction}. Recall that $\rho $ is a lift to
$A_{PGL_3}^{*}$ of
\[
\left( \alpha c_3\left( W\right) \right) _{\mid Diag_{\mathrm{sl}_3}^{*}}\in
\text{ }_3\left( A_{A_3\ltimes T}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right)
\right) ^{C_2}\simeq \text{ }_3A_{\Gamma _3}^{*}\left( Diag_{\mathrm{sl}%
_3}^{*}\right) \simeq \text{ }_3A_{PGL_3}^{*}\left( U\right) ,
\]
with $\alpha c_3\left( W\right) \in $ $_3A_{A_3\ltimes T}^{*}$. To prove (I)
it is then enough to show that
\[
\mathrm{cl}\left( \alpha c_3\left( W\right) \right) _{\mid Diag_{\mathrm{sl}%
_3}^{*}}\neq 0\text{ in }H_{A_3\ltimes T}^8\left( Diag_{\mathrm{sl}_3}^{*},%
\mathbf{Z}\right) \text{.}
\]
If $A_3\times \mathbf{\mu }_3\hookrightarrow A_3\ltimes T$, we will get this
by showing
\begin{equation}
\mathrm{cl}\left( \alpha c_3\left( W\right) \right) _{\mid Diag_{\mathrm{sl}%
_3}^{*}}\neq 0\text{ in }H_{A_3\times \mathbf{\mu }_3}^8\left( Diag_{\mathrm{%
sl}_3}^{*},\mathbf{Z}\right) \label{pippo}
\end{equation}
(writing again $\mathrm{cl}\left( \alpha c_3\left( W\right) \right) $ for
its the restriction to $H_{A_3\times \mathbf{\mu }_3}^8$). Let us consider
the localization exact sequences for cohomology, corresponding to $Diag_{%
\mathrm{sl}_3}\supset Diag_{\mathrm{sl}_3}\backslash \left\{ 0\right\}
\supset Diag_{\mathrm{sl}_3}^{*}$
\begin{equation}
H_{A_3\times \mathbf{\mu }_3}^4\stackrel{\cdot (-\alpha ^2)}{\longrightarrow
}H_{A_3\times \mathbf{\mu }_3}^8\left( Diag_{\mathrm{sl}_3},\mathbf{Z}%
\right) \stackrel{p}{\longrightarrow }H_{A_3\times \mathbf{\mu }_3}^8\left(
Diag_{\mathrm{sl}_3}\smallsetminus \left\{ 0\right\} ,\mathbf{Z}\right)
\label{cinque.uno}
\end{equation}
\begin{equation}
H_{\mathbf{\mu }_3}^6\simeq H_{A_3\times \mathbf{\mu }_3}^6\stackrel{}{%
\left( Z,\mathbf{Z}\right) \stackrel{i_{*}}{\longrightarrow }}H_{A_3\times
\mathbf{\mu }_3}^8\left( Diag_{\mathrm{sl}_3}\smallsetminus \left\{
0\right\} ,\mathbf{Z}\right) \stackrel{q}{\longrightarrow }H_{A_3\times
\mathbf{\mu }_3}^8\left( Diag_{\mathrm{sl}_3}^{*},\mathbf{Z}\right)
\label{cinque.due}
\end{equation}
where $i:Z\doteq \left( Diag_{\mathrm{sl}_3}\backslash \left\{ 0\right\}
\right) \backslash Diag_{\mathrm{sl}_3}^{*}\hookrightarrow Diag_{\mathrm{sl}%
_3}\backslash \left\{ 0\right\} $ and we used that $Z\simeq A_3\times
\mathbf{C}^{*}$, $A_3\ltimes T$-equivariantly. If $\mathbf{C}_{\chi ,\mathbf{%
\mu }_3}$ (respectively, $\mathbf{C}_{perm,A_3}^3$) denotes the $\mathbf{\mu
}_3$-representation given by multiplication by the character $\chi =\exp
\left( i2\pi /3\right) $ (respectively, the $A_3$-permutation
representation), we have $W\simeq \mathbf{C}_{\chi ,\mathbf{\mu }%
_3}\boxtimes \mathbf{C}_{perm,A_3}^3$ as $A_3\times \mathbf{\mu }_3$%
-representations. Then, if we let $H_{\mathbf{\mu }_3}^{*}=\mathbf{Z}\left[
\beta \right] /\left( 3\beta \right) $, $H_{A_3}^{*}=\mathbf{Z}\left[ \alpha
\right] /\left( 3\alpha \right) $, the Chern roots of $W$ are $\left\{ \beta
+\alpha ,\beta -\alpha ,\beta \right\} $ and
\[
\mathrm{cl}\left( \alpha c_3\left( W\right) \right) =\left( \beta ^2-\alpha
^2\right) \alpha \beta \in H_{A_3\times \mathbf{\mu }_3}^8\text{.}
\]
Now we claim $i_{*}=0$ in (\ref{cinque.due}). In fact, consider the pullback
$E$ of $\mathbf{C}_{\chi ,\mathbf{\mu }_3}$ to $Diag_{\mathrm{sl}%
_3}\backslash \left\{ 0\right\} $ and view $E$ as an $A_3\times \mathbf{\mu }%
_3$-equivariant vector bundle on $Diag_{\mathrm{sl}_3}\backslash \left\{
0\right\} $, with $A_3$ acting trivially on $E$. Obviously, $i^{*}\left(
c_1\left( E\right) \right) =\beta $. But we also have $i_{*}\left( 1\right)
=0$ since
\[
Z=D^{-1}\left( \left\{ 0\right\} \right) ,
\]
where
\begin{eqnarray}
D &:&Diag_{\mathrm{sl}_3}\backslash \left\{ 0\right\} \rightarrow \mathbf{A}%
^1 \nonumber \\
\left( \lambda _1,\lambda _2,\lambda _3\right) &\longmapsto &\left( \lambda
_1-\lambda _2\right) \left( \lambda _1-\lambda _3\right) \left( \lambda
_2-\lambda _3\right)
\end{eqnarray}
is the square root of the discriminant (which is $A_3\times \mathbf{\mu }_3$%
-equivariant!). By projection formula, $i_{*}=0$ and $q$ is injective.
So, we are left to show that $p\left( \left( \beta ^2-\alpha ^2\right)
\alpha \beta \right) =p\left( \alpha \beta ^3\right) \neq 0$ in (\ref
{cinque.uno}). Now observe that
\[
H_{A_3\times \mathbf{\mu }_3}^{2n}\simeq \left( H_{A_3}^{*}\otimes H_{%
\mathbf{\mu }_3}^{*}\right) ^{2n}
\]
by K\"unneth formula, since
\[
\bigoplus_{p+q=2n+1}Tor_1^{\mathbf{Z}}\left( H_{A_3}^p,H_{\mathbf{\mu }%
_3}^q\right) =0
\]
(either $p$ or $q$ being odd in every summand). So
\begin{eqnarray*}
H_{A_3\times \mathbf{\mu }_3}^8 &=&\mathbf{Z}/3\left\langle \alpha ^4,\alpha
^3\beta ,\alpha ^2\beta ^2,\alpha \beta ^3,\beta ^4\right\rangle \\
H_{A_3\times \mathbf{\mu }_3}^4 &=&\mathbf{Z}/3\left\langle \alpha ^2,\alpha
\beta ,\beta ^2\right\rangle
\end{eqnarray*}
and $\alpha \beta ^3\notin im\left( \cdot (-\alpha ^2)\right) $ i.e. $%
p\left( \alpha \beta ^3\right) \neq 0$.
(II) Consider the Leray spectral sequence
\[
E_2^{pq}=H^p\left( \mathrm{B}PGL_3,H^q\left( \mathrm{B}\mathbf{\mu }_3,%
\mathbf{Z}\right) \right) \Rightarrow H^{p+q}\left( \mathrm{B}SL_3,\mathbf{Z}%
\right) .
\]
By Lemma \ref{cohom}, its (first quadrant) $E_2$-term\footnote{%
We write only the parts we'll need.} is:
\[
\begin{tabular}{|cccccccccccc|}
\hline
\multicolumn{1}{|c|}{$\cdots $} & \multicolumn{1}{c|}{} &
\multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} &
\multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} &
\multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & $%
\vdots $ \\ \hline
\multicolumn{1}{|c|}{$\alpha ^4$} & \multicolumn{1}{c|}{$0$} &
\multicolumn{1}{c|}{$\alpha ^3\xi _2$} & \multicolumn{1}{c|}{$\alpha ^4y_3$}
& \multicolumn{1}{c|}{$\alpha ^4y_4$} & \multicolumn{1}{c|}{$0$} &
\multicolumn{1}{c|}{$\alpha ^4y_6$} & \multicolumn{1}{c|}{$\alpha ^3\xi _7$}
& \multicolumn{1}{c|}{$\alpha ^4x_8$, $\alpha ^4y_8$} & \multicolumn{1}{c|}{}
& \multicolumn{1}{c|}{} & \\ \hline
\multicolumn{1}{|c|}{$0$} & \multicolumn{1}{c|}{$0$} & \multicolumn{1}{c|}{$%
0 $} & \multicolumn{1}{c|}{$0$} & \multicolumn{1}{c|}{$0$} &
\multicolumn{1}{c|}{$0$} & \multicolumn{1}{c|}{$0$} & \multicolumn{1}{c|}{$0$%
} & \multicolumn{1}{c|}{$0$} & \multicolumn{1}{c|}{$0$} &
\multicolumn{1}{c|}{$0$} & \\ \hline
\multicolumn{1}{|c|}{$\alpha ^3$} & \multicolumn{1}{c|}{$0$} &
\multicolumn{1}{c|}{$\alpha ^2\xi _2$} & \multicolumn{1}{c|}{$\alpha ^3y_3$}
& \multicolumn{1}{c|}{$\alpha ^3y_4$} & \multicolumn{1}{c|}{$0$} &
\multicolumn{1}{c|}{$\alpha ^3y_6$} & \multicolumn{1}{c|}{$\alpha ^2\xi _7$}
& \multicolumn{1}{c|}{$\alpha ^3x_8$, $\alpha ^3y_8$} & \multicolumn{1}{c|}{}
& \multicolumn{1}{c|}{} & \\ \hline
\multicolumn{1}{|c|}{$0$} & \multicolumn{1}{c|}{$0$} & \multicolumn{1}{c|}{$%
0 $} & \multicolumn{1}{c|}{$0$} & \multicolumn{1}{c|}{$0$} &
\multicolumn{1}{c|}{$0$} & \multicolumn{1}{c|}{$0$} & \multicolumn{1}{c|}{$0$%
} & \multicolumn{1}{c|}{$0$} & \multicolumn{1}{c|}{$0$} &
\multicolumn{1}{c|}{$0$} & \\ \hline
\multicolumn{1}{|c|}{$\alpha ^2$} & \multicolumn{1}{c|}{$0$} &
\multicolumn{1}{c|}{$\alpha \xi _2$} & \multicolumn{1}{c|}{$\alpha ^2y_3$} &
\multicolumn{1}{c|}{$\alpha ^2y_4$} & \multicolumn{1}{c|}{$0$} &
\multicolumn{1}{c|}{$\alpha ^2y_6$} & \multicolumn{1}{c|}{$\alpha \xi _7$} &
\multicolumn{1}{c|}{$\alpha ^2x_8$, $\alpha ^2y_8$} & \multicolumn{1}{c|}{}
& \multicolumn{1}{c|}{} & \\ \hline
\multicolumn{1}{|c|}{$0$} & \multicolumn{1}{c|}{$0$} & \multicolumn{1}{c|}{$%
0 $} & \multicolumn{1}{c|}{$0$} & \multicolumn{1}{c|}{$0$} &
\multicolumn{1}{c|}{$0$} & \multicolumn{1}{c|}{$0$} & \multicolumn{1}{c|}{$0$%
} & \multicolumn{1}{c|}{$0$} & \multicolumn{1}{c|}{$0$} &
\multicolumn{1}{c|}{$0$} & \\ \hline
\multicolumn{1}{|c|}{$\alpha $} & \multicolumn{1}{c|}{$0$} &
\multicolumn{1}{c|}{$\xi _2$} & \multicolumn{1}{c|}{$\alpha y_3$} &
\multicolumn{1}{c|}{$\alpha y_4$} & \multicolumn{1}{c|}{$0$} &
\multicolumn{1}{c|}{$\alpha y_6$} & \multicolumn{1}{c|}{$\xi _7$} &
\multicolumn{1}{c|}{$\alpha x_8$, $\alpha y_8$} & \multicolumn{1}{c|}{} &
\multicolumn{1}{c|}{} & \\ \hline
\multicolumn{1}{|c|}{$0$} & \multicolumn{1}{c|}{$0$} & \multicolumn{1}{c|}{$%
0 $} & \multicolumn{1}{c|}{$0$} & \multicolumn{1}{c|}{$0$} &
\multicolumn{1}{c|}{$0$} & \multicolumn{1}{c|}{$0$} & \multicolumn{1}{c|}{$0$%
} & \multicolumn{1}{c|}{$0$} & \multicolumn{1}{c|}{$0$} &
\multicolumn{1}{c|}{$0$} & \\ \hline
\multicolumn{1}{|c|}{$\mathbf{Z}$} & \multicolumn{1}{c|}{$0$} &
\multicolumn{1}{c|}{$0$} & \multicolumn{1}{c|}{$y_3\mathbf{Z}/3$} &
\multicolumn{1}{c|}{$y_4\mathbf{Z}/3$} & \multicolumn{1}{c|}{$0$} &
\multicolumn{1}{c|}{$y_6\mathbf{Z}$} & \multicolumn{1}{c|}{$0$} &
\multicolumn{1}{c|}{$x_8\mathbf{Z}\oplus y_8\mathbf{Z}/3$} &
\multicolumn{1}{c|}{$0$} & \multicolumn{1}{c|}{$\mathbf{Z}$} & $\cdots $ \\
\hline
\end{tabular}
\]
where from the second row up, the coefficients are in $\mathbf{Z}/3$.
One of the edge maps is
\[
H^8\left( \mathrm{B}PGL_3,\mathbf{Z}\right) =E_2^{8,0}\twoheadrightarrow
E_\infty ^{8,0}=F^8H^8\left( \mathrm{B}SL_3,\mathbf{Z}\right)
\hookrightarrow H^8\left( \mathrm{B}SL_3,\mathbf{Z}\right)
\]
so we have to show that $F^8H^8\left( \mathrm{B}SL_3,\mathbf{Z}\right) \ $%
has index at least $9$ in
\[
H^8(\mathrm{B}SL_3,\mathbf{Z})\simeq \mathbf{Z}\cdot \alpha _2^2.
\]
First of all, note that $d_{(3)}(\alpha )=\pm y_3$ since
\[
E_\infty ^{3,0}=F^3H^3\left( \mathrm{B}SL_3,\mathbf{Z}\right)
\hookrightarrow H^3\left( \mathrm{B}SL_3,\mathbf{Z}\right) =0
\]
and both $\alpha $ and $y_3$ are $3$-torsion; we choose $y_3$ to have the
plus sign. Therefore
\[
d_{(3)}\left( \alpha ^2y_3\right) =2\alpha y_3^2+\alpha ^2d_{(3)}(y_3)=0
\]
since $y_3^2$ is $3$-torsion in $H^6\left( \mathrm{B}PGL_3,\mathbf{Z}\right)
\simeq \mathbf{Z}$, hence is zero.
Then
\[
E_2^{62}=E_3^{62}=E_4^{62}=E_\infty ^{62}\simeq \mathbf{Z}/3
\]
and we have the first $3$ factor of the desired index. Finally we have
\[
d_{(3)}\left( \alpha ^2y_4\right) =2\alpha y_3y_4+\alpha ^2d_{(3)}(y_4)=0
\]
since $y_3y_4\in H^7\left( \mathrm{B}PGL_3,\mathbf{Z}\right) =0$; then
\[
E_2^{44}=E_3^{44}=E_4^{44}=E_\infty ^{44}\simeq \mathbf{Z}/3
\]
yielding the other $3$ factor in the index of $F^8H^8\left( \mathrm{B}SL_3,%
\mathbf{Z}\right) \ $in $H^8(\mathrm{B}SL_3,\mathbf{Z})\simeq \mathbf{Z}%
\cdot \alpha _2^2$.
(III) As already observed, we have $c_2\left( \mathrm{sl}_3\right) ^2\mapsto
36\alpha _2^2$ via the pull back (use (I))
\[
\phi :\mathbf{Z}\oplus \mathrm{cl}\left( \rho \right) \cdot \left( \mathbf{Z/%
}3\right) \simeq H^8\left( \mathrm{B}PGL_3,\mathbf{Z}\right) \longrightarrow
H^8\left( \mathrm{B}SL_3,\mathbf{Z}\right) \simeq \mathbf{Z}\cdot \alpha
_2^2.
\]
whose kernel is $3$-torsion; combining this with (II), we get that the image
of $\phi $ has exactly index $9$. Therefore
\[
H^8\left( \mathrm{B}PGL_3,\mathbf{Z}_{(3)}\right) \simeq \mathbf{Z}%
_{(3)}\cdot j_{\bullet }\left( c_2\left( \mathrm{sl}_3\right) ^2\right)
\oplus \left( \mathbf{Z/}3\right) \cdot j_{\bullet }\left( \mathrm{cl}\left(
\rho \right) \right) ,
\]
where $j_{\bullet }:H^{*}(\mathrm{B}PGL_3,\mathbf{Z})\rightarrow H^{*}(%
\mathrm{B}PGL_3,\mathbf{Z}_{(3)})$ is the morphism induced by the
localization $j:\mathbf{Z}\rightarrow \mathbf{Z}_{(3)}$.
(IV) By \cite{KY} Cor. 4.7, we know that
\[
H^8\left( \mathrm{B}PGL_3,\mathbf{Z}/3\right) \simeq \left( \mathbf{Z}%
/3\right) \cdot y_2^4\oplus \left( \mathbf{Z/}3\right) \cdot y_8,
\]
and that the second generator $y_8$ is not in the Chern subring of $%
H^{*}\left( \mathrm{B}PGL_3,\mathbf{Z}/3\right) $. By the Bockstein exact
sequence, the natural map
\[
\left( j_{(3)}\right) _{\bullet }:H^8\left( \mathrm{B}PGL_3,\mathbf{Z}%
_{(3)}\right) \rightarrow H^8\left( \mathrm{B}PGL_3,\mathbf{Z}/3\right)
\]
is surjective since $H^9\left( \mathrm{B}PGL_3,\mathbf{Z}_{(3)}\right) =0$.
Therefore there exists an element $\xi =\alpha j_{\bullet }\left( c_2\left(
\mathrm{sl}_3\right) ^2\right) +\beta j_{\bullet }\left( \mathrm{cl}\left(
\rho \right) \right) \in H^8\left( \mathrm{B}PGL_3,\mathbf{Z}_{(3)}\right) $
such that $\left( j_{(3)}\right) _{\bullet }\left( \xi \right) =y_8$. In
particular, $\mathrm{cl}\left( \rho \right) $ cannot be in the Chern subring
of $H^{*}\left( \mathrm{B}PGL_3,\mathbf{Z}\right) $. \TeXButton{End Proof}
{\endproof}
\begin{remark}
For a different proof of Theorem \ref{result1}, which does not depend on
Kono-Yagita's results on $H^{*}\left( \mathrm{B}PGL_3,\mathbf{Z}/3\right) $
(and in fact does not depend on cohomology at all), see the Appendix.
\end{remark}
\begin{lemma}
\label{teta3tors} $\Theta _{1,1}^{\left( 2\right) },\Theta _{0,1}^{\left(
2\right) },\Theta _{1,0}^{\left( 3\right) },\Theta _{1,0}^{\left( 4\right)
},\Theta _{0,0}^{(4)},\Theta _{1,0}^{\left( 5\right) },\Theta
_{0,0}^{(5)},\Theta _{1,0}^{\left( 6\right) },\Theta _{1,0}^{\left( 7\right)
}$ and $\Theta _{1,0}^{\left( 8\right) }$ are $3$-torsion.
\end{lemma}
\TeXButton{Proof}{\proof} All the $\Theta $'s are supported on the
complement of $U$ and so they all go to zero via $A_{PGL_3}^{*}\rightarrow
A_T^{*}$, since this map factors through $A_{PGL_3}^{*}\rightarrow
A_{PGL_3}^{*}\left( U\right) $. But, by \cite{EG1}, Prop. 6, the rational
pullback
\[
A_{PGL_3}^{*}\otimes \mathbf{Q}\longrightarrow \left( A_T^{*}\right)
^{S_3}\otimes \mathbf{Q}
\]
is an isomorphism, so the $\Theta $'s are torsion and hence $3$-torsion by
Cor. \ref{cortorsion}. \TeXButton{End Proof}{\endproof}
\begin{remark}
Note that $\mathrm{cl}\left( \chi \right) =0$ since $\chi $ is torsion while
$H^{12}\left( \mathrm{B}PGL_3,\mathbf{Z}\right) $ is torsion free by lemma
\ref{cohom}.
\end{remark}
\begin{lemma}
\label{grado4}$\Theta _{1,0}^{\left( 4\right) }$ and $\Theta _{0,0}^{\left(
4\right) }$ are in the kernel of the cycle map $\mathrm{cl}%
:A_{PGL_3}^{*}\rightarrow H^{*}(\mathrm{B}PGL_3,\mathbf{Z})$.
\end{lemma}
\TeXButton{Proof}{\proof} By part (I) of the proof of Th. \ref{result1}, $%
\mathrm{cl}\left( \rho \right) $ generates the $3$-torsion of $H^8(\mathrm{B}%
PGL_3,\mathbf{Z})$ and moreover $\mathrm{cl}\left( \rho \right) _{\mid
U}\neq 0$ in $H_{PGL_3}^8\left( U,\mathbf{Z}\right) $, where $U\subset
\mathrm{sl}_3$ is the open subscheme of matrices with distinct eigenvalues.
Since $\Theta _{1,0}^{\left( 4\right) }$ and $\Theta _{0,0}^{\left( 4\right)
}$ are both $3$-torsion in $A_{PGL_3}^4$, we must have
\begin{eqnarray*}
\mathrm{cl}\left( \Theta _{1,0}^{\left( 4\right) }\right) &=&A\cdot \mathrm{%
cl}\left( \rho \right) \\
\mathrm{cl}\left( \Theta _{0,0}^{\left( 4\right) }\right) &=&B\cdot \mathrm{%
cl}\left( \rho \right) \text{ .}
\end{eqnarray*}
But $\Theta _{1,0}^{\left( 4\right) }$ and $\Theta _{0,0}^{\left( 4\right) }$
have supports in the complement of $U$, so $A=B=0$. \TeXButton{End Proof}
{\endproof}
\begin{remark}
\label{teta8}Note that also the generator $\Theta _{1,0}^{\left( 8\right) }$
can be chosen in such a way that
\[
\mathrm{cl}\left( \Theta _{1,0}^{\left( 8\right) }\right) =0.
\]
In fact $c_8\left( \mathrm{sl}_3\right) \neq 0$ in $H^{16}(\mathrm{B}PGL_3,%
\mathbf{Z})$ by \cite{KY}, Lemma 3.18 and
\[
H^{16}(\mathrm{B}PGL_3,\mathbf{Z})\simeq \mathbf{Z}\oplus \mathbf{Z}\oplus
\mathbf{Z}/3
\]
(Lemma. \ref{cohom}), therefore
\[
\mathrm{cl}\left( \Theta _{1,0}^{\left( 8\right) }\right) =Ac_8\left(
\mathrm{sl}_3\right) .
\]
Now observe that
\[
c_8\left( \mathrm{sl}_3\right) _{\mid \mathrm{sl}_3\backslash Z_0\cup
\left\{ 0\right\} }=0
\]
while $\Theta _{1,0}^{\left( 8\right) }$ is a lift of $\left( j_{1,0}\right)
_{*}\left( \lambda _1\lambda _2^2\right) $ where
\[
j_{1,0}:Z_{1,0}\hookrightarrow \mathrm{sl}_3\backslash Z_0\cup \left\{
0\right\} \text{ };
\]
thus we can choose a lift $\Theta _{1,0}^{\left( 8\right) }$ such that $A=0$.
\end{remark}
\begin{proposition}
\label{allzero}The elements
\[
\left\{ \Theta _{1,1}^{\left( 2\right) },\Theta _{0,1}^{\left( 2\right)
},\Theta _{1,0}^{\left( 3\right) },\Theta _{1,0}^{\left( 4\right) },\Theta
_{0,0}^{(4)},\Theta _{1,0}^{\left( 5\right) },\Theta _{0,0}^{(5)},\Theta
_{1,0}^{\left( 6\right) },\Theta _{1,0}^{\left( 7\right) },\Theta
_{1,0}^{\left( 8\right) }\right\}
\]
are all zero in $A_{PGL_3}^{*}$.
\end{proposition}
\TeXButton{Proof}{\proof}We first prove that
\begin{equation}
\Theta _{1,1}^{\left( 2\right) }=\Theta _{0,1}^{\left( 2\right) }=\Theta
_{1,0}^{\left( 3\right) }=\Theta _{1,0}^{\left( 4\right) }=\Theta
_{0,0}^{(4)}=\Theta _{1,0}^{\left( 5\right) }=\Theta _{0,0}^{(5)}=\Theta
_{1,0}^{(7)}=0\text{.} \label{firsthalf}
\end{equation}
Consider the commutative diagram
\[
\begin{tabular}{lll}
$A_{PGL_3}^{*}$ & $\stackrel{cl}{\longrightarrow }$ & $H^{*}\left( \mathrm{B}%
PGL_3,\mathbf{Z}\right) $ \\
$\downarrow $ & & $\downarrow $ \\
$A_{\Gamma _3}^{*}$ & $\stackunder{cl}{\longrightarrow }$ & $H^{*}\left(
\mathrm{B}\Gamma _3,\mathbf{Z}\right) $%
\end{tabular}
\]
where the vertical arrows are injective by Theorem \ref{gottliebtotaro}. We
know that
\[
\Theta _{1,1}^{\left( 2\right) },\Theta _{0,1}^{\left( 2\right) },\Theta
_{1,0}^{\left( 3\right) },\Theta _{1,0}^{\left( 4\right) },\Theta
_{0,0}^{(4)},\Theta _{1,0}^{\left( 5\right) },\Theta _{0,0}^{(5)},\Theta
_{1,0}^{(7)}
\]
are $3$-torsion and zero in cohomology (lemmas \ref{teta3tors}, \ref{cohom}
and \ref{grado4}), so (\ref{firsthalf}) will be proved if we show that
\[
_3cl:\text{ }_3A_{\Gamma _3}^{*}\longrightarrow \text{ }_3H^{*}\left(
\mathrm{B}\Gamma _3,\mathbf{Z}\right)
\]
is injective up to degree $5$ and in degree $7$. But, by the usual
transfer-trick, the restriction induces isomorphisms
\[
_3A_{\Gamma _3}^{*}\simeq \left( _3A_{A_3\ltimes T}^{*}\right) ^{C_2},\text{
}_3H^{*}\left( \mathrm{B}\Gamma _3,\mathbf{Z}\right) \simeq \left(
_3H^{*}\left( \mathrm{B}\left( A_3\ltimes T\right) ,\mathbf{Z}\right)
\right) ^{C_2}
\]
and it will be (more than) enough to show that
\[
cl:A_{A_3\ltimes T}^{*}\longrightarrow H^{*}\left( \mathrm{B}\left(
A_3\ltimes T\right) ,\mathbf{Z}\right)
\]
is injective up to degree $5$ and in degree $7$.
Recall (Prop. \ref{basico}) that $A_{A_3\ltimes T}^{*}$ is generated by
\begin{equation}
\left\{ \alpha ,c_2\left( W\right) ,c_3\left( W\right) ,\theta \doteq
\mathrm{tsf}_T^{A_3\ltimes T}\left( u_2^2u_3\right) \right\} \label{gen}
\end{equation}
where $W$ is the representation defined in (\ref{vudoppio}) and we identify $%
A_T^{*}$ with
\[
A_{T_{SL_3}}^{*}\simeq \mathbf{Z}\left[ u_1,u_2,u_3\right] \diagup \left(
u_1+u_2+u_3\right) ;
\]
moreover (see Lemma \ref{basico2}), we have
\begin{equation}
3\alpha =0,\text{ }\alpha \theta =0,\text{ }\alpha ^3+\alpha c_2\left(
W\right) =0, \label{rel}
\end{equation}
\[
3\left[ \left( 2\theta +3c_3\left( W\right) \right) ^2+4c_2\left( W\right)
^3+27c_3\left( W\right) ^2\right] =0.
\]
For the duration of this proof, we will denote $c_2\left( W\right) $ and $%
c_3\left( W\right) $ simply by $c_2$ and $c_3$; moreover, if $\xi \in
A_{A_3\ltimes T}^{*}$, we will write $\overline{\xi }$ for $cl\left( \xi
\right) $.
As shown in the proof of Th. \ref{result1}, we have
\[
H^{2n}\left( \mathrm{B}\left( A_3\times \mathbf{\mu }_3\right) ,\mathbf{Z}%
\right) \simeq \left( H^{*}\left( \mathrm{B}A_3,\mathbf{Z}\right) \otimes
\left( \mathrm{B}\mathbf{\mu }_3,\mathbf{Z}\right) \right) ^{2n}
\]
and
\begin{equation}
\overline{c_2\left( W\right) }_{\mid A_3\times \mathbf{\mu }_3}=-\overline{%
\alpha }^2,\text{ }\overline{c_3\left( W\right) }_{\mid A_3\times \mathbf{%
\mu }_3}=\overline{\beta }\left( \overline{\beta }^2-\overline{\alpha }%
^2\right) \label{a3mi3}
\end{equation}
where
\[
H^{*}\left( \mathrm{B}A_3,\mathbf{Z}\right) =\frac{\mathbf{Z}\left[
\overline{\alpha }\right] }{\left( \overline{\alpha }\right) },\text{ }%
H^{*}\left( \mathrm{B}\mathbf{\mu }_3,\mathbf{Z}\right) =\frac{\mathbf{Z}%
\left[ \overline{\beta }\right] }{\left( 3\overline{\beta }\right) }.
\]
In the following computations we will freely use that the cycle class map
respects Chern classes, restrictions and transfers and that
\[
cl:A_T^{*}\longrightarrow H^{*}\left( \mathrm{B}T,\mathbf{Z}\right)
\]
is an isomorphism.
If $\xi \in \ker cl\cap A_{A_3\ltimes T}^1,$ we have
\[
\xi =A\alpha \text{ \quad and \quad }A\overline{\alpha }=0
\]
for some $A\in \mathbf{Z}$; restricting this to $A_3$ (in cohomology) we
then get $A\equiv 0$ $\limfunc{mod}3$, hence $\xi =0$.
If $\xi \in \ker cl\cap A_{A_3\ltimes T}^2,$ we have
\[
\xi =A\alpha ^2+Bc_2,\text{ }A\overline{\alpha }^2+B\overline{c_2}=0
\]
for some $A,B\in \mathbf{Z}$; restricting to $T$, we get $B=0$ then,
restricting to $A_3$, we get $A\equiv 0$ $\limfunc{mod}3$. Therefore, $\xi
=0 $.
If $\xi \in \ker cl\cap A_{A_3\ltimes T}^3,$ we have
\[
\xi =A\alpha ^3+Bc_3+C\theta
\]
\[
A\overline{\alpha }^3+B\overline{c_3}+C\overline{\theta }=0
\]
for some $A,B,C\in \mathbf{Z}$; restricting to $T$, we get $B=C=0$ since $%
\overline{c_3}_{\mid T}$ and $\overline{\theta }_{\mid T}$ are linearly
independent $H^{*}\left( \mathrm{B}T,\mathbf{Z}\right) $. Restricting then
to $A_3$, we get $A\equiv 0$ $\limfunc{mod}3$, hence $\xi =0$.
If $\xi \in \ker cl\cap A_{A_3\ltimes T}^4,$ we have
\[
\xi =A\alpha ^4+B\alpha c_3+Cc_2^2
\]
\[
A\overline{\alpha }^4+B\overline{\alpha }\overline{c_3}+C\overline{c_2}^2=0
\]
for some $A,B,C\in \mathbf{Z}$; restricting to $T$, we get $C=0$.
Restricting then to $A_3\times \mathbf{\mu }_3$, from (\ref{a3mi3}) we get $%
B\equiv A\equiv 0$ $\limfunc{mod}3$, hence $\xi =0$.
If $\xi \in \ker cl\cap A_{A_3\ltimes T}^5,$ we have
\[
\xi =A\alpha ^5+B\alpha ^2c_3+Cc_2c_3
\]
\[
A\overline{\alpha }^5+B\overline{\alpha }^2\overline{c_3}+C\overline{c_2}%
\overline{c_3}=0
\]
for some $A,B,C\in \mathbf{Z}$; restricting to $T$, we get $C=0$.
Restricting then to $A_3\times \mathbf{\mu }_3$, from (\ref{a3mi3}) we get $%
B\equiv A\equiv 0$ $\limfunc{mod}3$, hence $\xi =0$.
Finally, if $\xi \in \ker cl\cap A_{A_3\ltimes T}^7,$ we have
\[
\xi =A\alpha ^7+B\alpha ^4c_3+Cc_2^2c_3+Dc_2^2\theta +E\alpha c_3^2
\]
\[
A\overline{\alpha }^7+B\overline{\alpha }^4\overline{c_3}+C\overline{c_2}^2%
\overline{c_3}+D\overline{c_2}^2\overline{\theta }+E\overline{\alpha }%
\overline{c_3}^2=0
\]
for some $A,B,C,D,E\in \mathbf{Z}$; restricting to $T$, we get $C=D=0$ since
$\overline{c_2}_{\mid T}\neq 0$ and $\left( \overline{c_3}_{\mid T},%
\overline{\theta }_{\mid T}\right) $ are linearly independent in the domain $%
H^{*}\left( \mathrm{B}T,\mathbf{Z}\right) $. Restricting then to $A_3\times
\mathbf{\mu }_3$, from (\ref{a3mi3}) we get $A\equiv B\equiv E\equiv 0$ $%
\limfunc{mod}3$, hence $\xi =0$. This concludes the proof of (\ref{firsthalf}%
).
Now we prove the remaining relations
\begin{equation}
\Theta _{1,0}^{\left( 6\right) }=\Theta _{1,0}^{\left( 8\right) }=0.
\label{secondhalf}
\end{equation}
First observe that $\Theta _{1,0}^{\left( 6\right) }$ and $\Theta
_{1,0}^{\left( 8\right) }$ are $3$-torsion and zero in cohomology (with $%
\Theta _{1,0}^{\left( 8\right) }$ chosen as in Remark \ref{teta8}). Since
they are lifts of elements having supports in the complement of $U\subset
\mathrm{sl}_3$, their restrictions to $A_{\Gamma _3}^{*}$ are in the kernel
of
\[
A_{\Gamma _3}^{*}\longrightarrow A_{\Gamma _3}^{*}\left( Diag_{\mathrm{sl}%
_3}^{*}\right) \simeq A_{PGL_3}^{*}\left( U\right) \text{.}
\]
and in particular:
\[
\left\{ \Theta _{1,0\mid A_3\ltimes T}^{\left( 6\right) },\Theta _{1,0\mid
A_3\ltimes T}^{\left( 8\right) }\right\} \subset \ker \left( g:A_{A_3\ltimes
T}^{*}\longrightarrow A_{A_3\ltimes T}^{*}\left( Diag_{\mathrm{sl}%
_3}^{*}\right) \right) .
\]
By Lemma \ref{basico2} (ii) and (\ref{gen}), (\ref{rel}), we must have
\begin{eqnarray*}
\Theta _{1,0\mid A_3\ltimes T}^{\left( 6\right) } &=&\alpha ^2\left( A\alpha
^4+B\alpha c_3+Cc_2^2\right) \\
\Theta _{1,0\mid A_3\ltimes T}^{\left( 8\right) } &=&\alpha ^2\left( D\alpha
^6+Ec_2^3+Fc_3^2+G\alpha ^3c_3\right)
\end{eqnarray*}
for some $A,...,E\in \mathbf{Z}$. Using again (\ref{rel}), we get
\begin{eqnarray*}
\Theta _{1,0\mid A_3\ltimes T}^{\left( 6\right) } &=&A^{\prime }\alpha
^6+B\alpha ^3c_3 \\
\Theta _{1,0\mid A_3\ltimes T}^{\left( 8\right) } &=&D^{\prime }\alpha
^6+Fc_3^2+G\alpha ^3c_3
\end{eqnarray*}
for some $A^{\prime },B,D^{\prime },F,G\in \mathbf{Z}$. Again denoting $%
cl\left( \xi \right) $ by $\overline{\xi }$ for $\xi \in A_{A_3\ltimes
T}^{*} $, we have
\begin{eqnarray*}
0 &=&\overline{\Theta _{1,0\mid A_3\ltimes T}^{\left( 6\right) }}=A^{\prime }%
\overline{\alpha }^6+B\overline{\alpha }^3\overline{c_3} \\
0 &=&\overline{\Theta _{1,0\mid A_3\ltimes T}^{\left( 8\right) }}=D^{\prime }%
\overline{\alpha }^6+F\overline{c_3}^2+G\overline{\alpha }^3\overline{c_3}
\end{eqnarray*}
in $H^{*}\left( \mathrm{B}\left( A_3\ltimes T\right) ,\mathbf{Z}\right) $.
Restricting these relations to $A_3\times \mathbf{\mu }_3$, by (\ref{a3mi3})
we obtain:
\begin{eqnarray*}
A^{\prime } &\equiv &B\equiv 0\text{ }\limfunc{mod}3 \\
D^{\prime } &\equiv &F\equiv G\equiv 0\text{ }\limfunc{mod}3
\end{eqnarray*}
i.e.
\begin{equation}
\Theta _{1,0\mid A_3\ltimes T}^{\left( 6\right) }=\text{ }\Theta _{1,0\mid
A_3\ltimes T}^{\left( 8\right) }=0 \label{qfinito}
\end{equation}
in $A_{A_3\ltimes T}^{*}$. But the restriction map induces an isomorphism
\[
_3A_{\Gamma _3}^{*}\simeq \left( _3A_{A_3\ltimes T}^{*}\right) ^{C_2}
\]
and then, we also get
\[
\Theta _{1,0\mid \Gamma _3}^{\left( 6\right) }=\text{ }\Theta _{1,0\mid
\Gamma _3}^{\left( 8\right) }=0
\]
in $A_{\Gamma _3}^{*}$. By Theorem \ref{gottliebtotaro} we finally get (\ref
{secondhalf}). \TeXButton{End Proof}{\endproof}
Thus we can summarize the main result obtained so far in the following:
\begin{theorem}
\label{generators}With the notation of (\ref{laststep}), $A_{PGL_3}^{*}$ is
generated by
\begin{equation}
\left\{ 2c_2\left( \mathrm{sl}_3\right) -c_2\left( Sym^3E\right) ,c_3\left(
Sym^3E\right) ,\rho ,\chi ,c_6\left( \mathrm{sl}_3\right) ,c_8\left( \mathrm{%
sl}_3\right) \right\} \label{generatori}
\end{equation}
where\emph{\ }$\deg \rho =4$\emph{, }$\deg \chi =6$.
\end{theorem}
\begin{remark}
We point out that
\[
2c_2\left( \mathrm{sl}_3\right) -c_2\left( Sym^3E\right) ,c_3\left(
Sym^3E\right) ,c_6\left( \mathrm{sl}_3\right) ,c_8\left( \mathrm{sl}%
_3\right)
\]
are nonzero (by checking their images in $A_{SL_3}^{*}$ or in cohomology)
and we will show in the next section that $\rho \neq 0$. Unfortunately, we
do not know whether $\chi $ is zero or not.
Note also that the generators $\rho $ and $\chi $, defined originally as
lifts from the open subset $U$ (therefore not unique \emph{a priori}) are
indeed uniquely defined since they have degrees $<8$ and $c_8\left( \mathrm{%
sl}_3\right) $ is the only generator coming from the complement of $U$.
\end{remark}
\section{Other relations and results on the cycle maps\ }
With the notations established in the preceding sections we have:
\begin{proposition}
\label{relations} The following relations hold among the generators of $%
A_{PGL_3}^{*}$:
\[
3\rho =3\chi =3c_8\left( \mathrm{sl}_3\right) =0
\]
\[
3\left( 27c_6\left( \mathrm{sl}_3\right) -c_3\left( Sym^3E\right)
^2-4\lambda ^3\right) =0
\]
\[
\rho ^2=c_8\left( \mathrm{sl}_3\right) .
\]
\end{proposition}
\TeXButton{Proof}{\proof}The pullback $\varphi :A_{PGL_3}^{*}\rightarrow
A_T^{*}$ factors through the composition
\[
\pi :A_{PGL_3}^{*}\twoheadrightarrow A_{PGL_3}^{*}\left( U\right) \simeq
A_{\Gamma _3}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right) \twoheadrightarrow
A_T^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right) ^{S_3}=\left( A_T^{*}\right)
^{S_3}\text{, }
\]
and, by definition of $\chi $ and $\rho $, $\pi \left( \chi \right) =\pi
\left( \rho \right) =0$. Since (\cite{EG1}, Prop. 6) the rational pullback $%
\varphi _{\mathbf{Q}}$ is an isomorphism, $\chi $ and $\rho $ are torsion
and then $3$-torsion by Cor. \ref{cortorsion}.
Since $\mathrm{sl}_3=E\otimes E^{\vee }-\mathbf{1}$, as $SL_3$%
-representations ($E$ being the standard representation), $c_8\left( \mathrm{%
sl}_3\right) $ is in the kernel of $A_{PGL_3}^{*}\rightarrow A_{SL_3}^{*}$,
so it is $3$-torsion (Prop. \ref{torsion}).
A long but straightforward computation\footnote{%
The basic fact here is that $c_6\left( \mathrm{sl}_3\right) $ restricts to
minus the discriminant, $4\alpha _2^3+27\alpha _3^2$, in $A_{SL_3}^{*}=%
\mathbf{Z}\left[ \alpha _2,\alpha _3\right] $, where $\alpha _i=c_i\left(
E\right) $.} shows that
\[
27c_6\left( \mathrm{sl}_3\right) -c_3\left( Sym^3E\right) ^2-4\lambda ^3\in
\ker \left( A_{PGL_3}^{*}\rightarrow A_{SL_3}^{*}\right)
\]
so that this element is $3$-torsion (again by Prop. \ref{torsion}).
By definition of $\rho $ and Lemma \ref{basico2}, we have
\begin{equation}
\text{ }\rho _{\mid A_3\ltimes T}=\alpha c_3\left( W\right) +A\alpha ^4
\label{ro}
\end{equation}
for some $A\in \mathbf{Z}/3$. Since
\[
_3A_{\Gamma _3}^{*}\simeq \left( _{\text{ }3}A_{A_3\ltimes T}^{*}\right)
^{C_2},
\]
by \label{lastrelation}Lemma \ref{basico2} (ii), $\rho ^2$ belongs to the
kernel of
\[
A_{PGL_3}^{*}\longrightarrow A_{PGL_3}^{*}\left( U\right) \simeq A_{\Gamma
_3}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right) .
\]
Therefore, since by Proposition \ref{allzero} all the generators of $%
A_{PGL_3}^{*}$ coming from the complement of $U$ are zero except for $%
c_8\left( \mathrm{sl}_3\right) $, we have
\begin{equation}
\rho ^2=Bc_8\left( \mathrm{sl}_3\right) \label{roquadro}
\end{equation}
for some $B\in \mathbf{Z}/3$.
Let us determine $A$ and $B$. Since $c_8\left( \mathrm{sl}_3\right) _{\mid
A_3}=0$, from (\ref{ro}) and (\ref{roquadro}), we get $A=0$ i.e.
\begin{equation}
\rho _{\mid A_3\ltimes T}=\alpha c_3\left( W\right) . \label{ro'}
\end{equation}
Straightforward computations show that
\[
c_8\left( \mathrm{sl}_3\right) _{\mid A_3\times \mathbf{\mu }_3}=\alpha
^2\beta ^2\left( \beta ^2-\alpha ^2\right) ^2
\]
\[
c_3\left( W\right) _{\mid A_3\times \mathbf{\mu }_3}=\beta \left( \beta
^2-\alpha ^2\right)
\]
in $A_{_{A_3\times \mathbf{\mu }_3}}^{*}=A_{A_3}^{*}\otimes A_{\mathbf{\mu }%
_3}^{*}=\mathbf{Z}\left[ \alpha \right] /\left( 3\alpha \right) \otimes
\mathbf{Z}\left[ \beta \right] /\left( 3\beta \right) $ and then (\ref
{roquadro}) and (\ref{ro'}) prove that $B=1$. \TeXButton{End Proof}
{\endproof}
We define the graded ring
\[
R^{*}\doteq \frac{\mathbf{Z}\left[ \lambda ,c_3\left( Sym^3E\right) ,\rho
,\chi ,c_6\left( \mathrm{sl}_3\right) ,c_8\left( \mathrm{sl}_3\right)
\right] }{\frak{R}}
\]
where
\[
\frak{R}\doteq \left( 3\rho ,3\chi ,3c_8\left( \mathrm{sl}_3\right) ,3\left(
27c_6\left( \mathrm{sl}_3\right) -c_3\left( Sym^3E\right) ^2-4\lambda
^3\right) ,\rho ^2-c_8\left( \mathrm{sl}_3\right) \right)
\]
and\emph{\ }$\deg \rho =4$,\emph{\ }$\deg \chi =6$.
This is our candidate for $A_{PGL_3}^{*}$. What we do know is that the
canonical morphism
\[
\pi :R^{*}\longrightarrow A_{PGL_3}^{*}
\]
is surjective (Th. \ref{generators}).
\begin{remark}
Note that it is immediately clear that $\pi _{\mathbf{Q}}:R^{*}\otimes
\mathbf{Q}\rightarrow A_{PGL_3}^{*}\otimes \mathbf{Q}$ is an isomorphism. In
fact
\[
R^{*}\otimes \mathbf{Q}=\frac{\mathbf{Q}\left[ \lambda ,c_3\left(
Sym^3E\right) ,c_6\left( \mathrm{sl}_3\right) \right] }{\left( 27c_6\left(
\mathrm{sl}_3\right) -c_3\left( Sym^3E\right) ^2-4\lambda ^3\right) }=
\]
\[
=\mathbf{Q}\left[ \lambda ,c_3\left( Sym^3E\right) \right] .
\]
Moreover, $\lambda \mapsto 3\alpha _2$ and $c_3\left( Sym^3E\right) \mapsto
27\alpha _3$ via
\[
A_{PGL_3}^{*}\longrightarrow A_{SL_3}^{*}=\mathbf{Z}\left[ \alpha _2,\alpha
_3\right]
\]
which is rationally an isomorphism (Prop. \ref{rational}). We will prove in
Proposition \ref{result0} (ii) that more is true: $R^{*}$ and $A_{PGL_3}^{*}$
are isomorphic after inverting $3$.
\end{remark}
We will now establish some properties of the cycle map
\[
\mathrm{cl}:A_{PGL_3}^{*}\longrightarrow H^{*}\left( \mathrm{B}PGL_3,\mathbf{%
Z}\right)
\]
and of Totaro's refined cycle map
\[
\widetilde{\mathrm{cl}}:A_{PGL_3}^{*}\longrightarrow MU^{*}\left( \mathrm{B}%
PGL_3\right) \otimes _{MU^{*}}\mathbf{Z.}
\]
\begin{remark}
\label{totaro}In \cite{KY} Kono and Yagita proved that in the
Atiyah-Hirzebruch spectral sequence for Brown-Peterson cohomology at the
prime $3$ (\cite{W})
\[
E_2^{pq}=H^p\left( \mathrm{B}PGL_3,BP^q\right) \Longrightarrow
BP^{p+q}\left( \mathrm{B}PGL_3\right)
\]
the $E_\infty $-term is generated as a $BP^{*}$-module by the top row i.e.
by
\[
im\left( BP^{*}\left( \mathrm{B}PGL_3\right) \longrightarrow H^{*}\left(
\mathrm{B}PGL_3,\mathbf{Z}_{\left( 3\right) }\right) \right) .
\]
As a consequence, the natural map
\[
\underline{\mathrm{cl}}:MU^{*}\left( \mathrm{B}PGL_3\right) \otimes _{MU^{*}}%
\mathbf{Z}\longrightarrow H^{*}\left( \mathrm{B}PGL_3,\mathbf{Z}\right)
\]
is injective.
\end{remark}
We have the following result\footnote{%
A stronger version of (i) will be proved in Theorem \ref{riassunto}.}:
\begin{proposition}
(i) \label{result0}$\mathrm{cl}$ and $\widetilde{\mathrm{cl}}$ are injective
after inverting $3$;
(ii) $\pi $ is an isomorphism after inverting $3$.
\end{proposition}
\TeXButton{Proof}{\proof} (i) $A_{PGL_3}^{*}$ has only $3$-torsion and $\ker
\mathrm{cl}$ is torsion (Section 2). Therefore $\mathrm{cl}=\underline{%
\mathrm{cl}}\circ \widetilde{\mathrm{cl}}$ is injective after inverting $3$
and the same is true for $\widetilde{\mathrm{cl}}$.
(ii) It is enough to prove that for any prime $p\neq 3$, the composition%
\footnote{$\left( \cdot \right) _{\left( p\right) }$ denotes localization at
the prime $p$.}
\[
\left( R^{*}\right) _{\left( p\right) }\stackrel{\pi _{(p)}}{\longrightarrow
}\left( A_{PGL_3}^{*}\right) _{\left( p\right) }\stackrel{\mathrm{cl}_{(p)}}{%
\longrightarrow }H^{*}(\mathrm{B}PGL_3,\mathbf{Z}_{\left( p\right) })
\]
is injective. Leray spectral sequence with $\mathbf{Z}_{\left( p\right) }$%
-coefficients:
\[
E_2^{pq}=H^p\left( \mathrm{B}PGL_3,H^q\left( \mathrm{B}\mathbf{\mu }_3,%
\mathbf{Z}_{\left( p\right) }\right) \right) \Longrightarrow H^{p+q}\left(
\mathrm{B}SL_3,\mathbf{Z}_{\left( p\right) }\right)
\]
collapses at the $E_2$-term since $H^{*}\left( \mathrm{B}\mathbf{\mu }_3,%
\mathbf{Z}_{\left( p\right) }\right) =\mathbf{Z}_{\left( p\right) }$,
concentrated in degree zero, thus yielding an ''edge'' isomorphism
(coinciding with the pullback):
\[
\varphi _{\left( p\right) }:H^{*}\left( \mathrm{B}PGL_3,\mathbf{Z}_{\left(
p\right) }\right) \simeq H^{*}\left( \mathrm{B}SL_3,\mathbf{Z}_{\left(
p\right) }\right) =\mathbf{Z}_{\left( p\right) }\left[ \alpha _2,\alpha
_3\right] \text{.}
\]
Now, consider the commutative diagram
\begin{equation}
\begin{tabular}{ccccc}
$\left( R^{*}\right) _{\left( p\right) }$ & $\stackrel{\pi _{\left( p\right)
}}{\longrightarrow }$ & $\left( A_{PGL_3}^{*}\right) _{\left( p\right) }$ & $%
\stackrel{\mathrm{cl}_{\left( p\right) }}{\longrightarrow }$ & $H^{*}\left(
\mathrm{B}PGL_3,\mathbf{Z}_{\left( p\right) }\right) $ \\
& & $^{\phi _{\left( p\right) }}\downarrow $ & & $\downarrow ^{\varphi
_{\left( p\right) }}$ \\
& & $\left( A_{SL_3}^{*}\right) _{\left( p\right) }$ & $\stackrel{}{%
\stackunder{\mathrm{cl}_{SL_3,\left( p\right) }}{\widetilde{\longrightarrow }%
}}$ & $H^{*}\left( \mathrm{B}SL_3,\mathbf{Z}_{\left( p\right) }\right) $%
\end{tabular}
\label{comm}
\end{equation}
and observe that for $p\neq 3$,
\[
\left( R^{*}\right) _{\left( p\right) }=\frac{\mathbf{Z}_{\left( p\right)
}\left[ \lambda ,c_3\left( Sym^3E\right) ,c_6\left( \mathrm{sl}_3\right)
\right] }{\left( 27c_6\left( \mathrm{sl}_3\right) -c_3\left( Sym^3E\right)
^2-4\lambda ^3\right) }=\mathbf{Z}_{\left( p\right) }\left[ \lambda
,c_3\left( Sym^3E\right) \right] .
\]
Since, as we already computed, $\phi \circ \pi \left( \lambda \right)
=3\alpha _2$, $\phi \circ \pi \left( c_3\left( Sym^3E\right) \right)
=27\alpha _3$, commutativity of (\ref{comm}) concludes the proof.
\TeXButton{End Proof}{\endproof}
The stronger result we can prove about $\widetilde{\mathrm{cl}}$ is the
following
\begin{theorem}
\label{riassunto}Totaro's refined cycle class map
\[
\widetilde{\mathrm{cl}}:A_{PGL_3}^{*}\longrightarrow MU^{*}(\mathrm{B}%
PGL_3)\otimes _{MU^{*}}\mathbf{Z}
\]
is surjective (and has $3$-torsion kernel).
\end{theorem}
\TeXButton{Proof}{\proof} $\ker \widetilde{\mathrm{cl}}$ is $3$-torsion
since it is torsion and $A_{PGL_3}^{*}$ has only $3$-torsion. So we are left
to prove surjectivity of $\widetilde{\mathrm{cl}}$. To do this, we first
prove that $\widetilde{\mathrm{cl}}$ is surjective (thus an isomorphism by
Prop. \ref{result0} (i)) after inverting $3$ and then that $\widetilde{%
\mathrm{cl}}$ is surjective when localized at the prime $3$.
$\underline{\mathrm{cl}}_{PGL_3}$ is an isomorphism after inverting $3$
since $H^{*}\left( \mathrm{B}PGL_3,\mathbf{Z}\left[ \frac 13\right] \right) $
is torsion free
(\cite{To1})\footnote{%
We briefly sketch the argument. Since the differentials in the
Atiyah-Hirzebruch spectral sequence
\[
F_2^{pq}=H^p\left( BPGL_3,MU^q\right) \Rightarrow MU^{p+q}\left(
BPGL_3\right)
\]
are always torsion, they must be $0$ if $3$ is inverted since there is only $%
3$-torsion (recall that $MU^{*}$ is torsion-free). Therefore $F_2^{pq}$
collapses when $3$ is inverted.}. So it is enough to prove that \textrm{cl}$%
_{PGL_3}$ is surjective when $3$ is inverted. Now, in the commutative
diagram
\[
\begin{tabular}{cc}
$A_{PGL_3}^{*}\left[ \frac 13\right] \stackrel{\mathrm{cl}_{PGL_3}\left[
\frac 13\right] }{\longrightarrow }$ & $H_{PGL_3}^{*}\left[ \frac 13\right] $
\\
$^\varphi \downarrow \quad $ & $^{\varphi ^{\prime }}\downarrow \quad $ \\
$A_{SL_3}^{*}\left[ \frac 13\right] \stackunder{\mathrm{cl}_{SL_3}\left[
\frac 13\right] }{\longrightarrow }$ & $H_{SL_3}^{*}\left[ \frac 13\right] $%
\end{tabular}
\]
$\varphi ^{\prime }$ is an isomorphism since the corresponding Leray
spectral sequence
\[
E_2^{pq}=H^p\left( \mathrm{B}PGL_3,H^q\left( \mathrm{B}\mathbf{\mu }_3,%
\mathbf{Z}\right) \right) \Rightarrow H^{p+q}\left( \mathrm{B}SL_3,\mathbf{Z}%
\right)
\]
collapses after inverting $3$, and \textrm{cl}$_{SL_3}$ is an isomorphism
even without inverting $3$. On the other hand, $\varphi $ is injective
because $\Phi :A_{PGL_3}^{*}\rightarrow A_{SL_3}^{*}$ has $3$-torsion kernel
and is surjective since
\begin{eqnarray*}
&&\ c_2\left( \mathrm{sl}_3\right) \stackrel{\Phi }{\longmapsto }6\alpha _2
\\
&&\ c_2\left( Sym^3E\right) \stackrel{\Phi }{\longmapsto }15\alpha _2 \\
&&\ c_3\left( Sym^3E\right) \stackrel{\Phi }{\longmapsto }27\alpha _3.
\end{eqnarray*}
Therefore $\mathrm{cl}_{PGL_3}\left[ \frac 13\right] $ is an isomorphism too.
So it remains to prove that the localization at the prime $3$%
\[
\left( \widetilde{\mathrm{cl}}_{PGL_3}\right) _{\left( 3\right) }:\left(
A_{PGL_3}^{*}\right) _{\left( 3\right) }\longrightarrow MU^{*}(\mathrm{B}%
PGL_3)\otimes _{MU^{*}}\mathbf{Z}_{\left( 3\right) }
\]
is surjective. By \cite{Q},
\[
MU^{*}(\mathrm{B}PGL_3)\otimes _{MU^{*}}\mathbf{Z}_{\left( 3\right) }\simeq
BP^{*}(\mathrm{B}PGL_3)\otimes _{BP^{*}}\mathbf{Z}_{\left( 3\right) }
\]
where $BP^{*}\left( X\right) $ denotes the Brown-Peterson cohomology of $X$
localized at the prime $3$ and
\[
BP^{*}=BP^{*}\left( pt\right) =\mathbf{Z}_{\left( 3\right) }\left[
v_1,\ldots ,v_n,\ldots \right] \longrightarrow \mathbf{Z}_{\left( 3\right) }
\]
($\deg v_i=-2\left( 3^i-1\right) $) sends each $v_i$ to zero (see also \cite
{W}). Kono and Yagita computed $BP^{*}(\mathrm{B}PGL_3)$ in \cite{KY}, Th.
4.9, as a $BP^{*}$-module; it is a quotient of the following $BP^{*}$-module
\[
\left( BP^{*}\mathbf{Z}_{\left( 3\right) }\left[ \left[ \widetilde{y_2}%
\right] \right] \widetilde{y_2}^2\oplus BP^{*}\oplus BP^{*}\mathbf{Z}%
_{\left( 3\right) }\left[ \left[ \widetilde{y_8}\right] \right] \widetilde{%
y_8}\right) \otimes \mathbf{Z}_{\left( 3\right) }\left[ \left[ \widetilde{%
y_{12}}\right] \right]
\]
and, if
\[
r:BP^{*}(\mathrm{B}PGL_3)\stackrel{s}{\longrightarrow }H^{*}\left( \mathrm{B}%
PGL_3,\mathbf{Z}_{\left( 3\right) }\right) \stackrel{j_{\bullet }}{%
\longrightarrow }H^{*}\left( \mathrm{B}PGL_3,\mathbf{Z}/3\right)
\]
(where $s$ is the natural map of generalized cohomology theories and $%
j_{\bullet }$ is induced by $j:\mathbf{Z}_{\left( 3\right) }\rightarrow
\mathbf{Z}/3$), $r$ has kernel $BP^{<0}\cdot BP^{*}(\mathrm{B}PGL_3)$ and
\begin{eqnarray*}
r\left( \widetilde{y_2}^2\right) &=&y_2^2\equiv c_2\left( \mathrm{sl}%
_3\right) \\
r\left( \widetilde{y_8}\right) &=&y_8^{} \\
r\left( \widetilde{y_{12}}\right) &=&y_{12}\equiv c_6\left( \mathrm{sl}%
_3\right) ,
\end{eqnarray*}
$y_8\in H^8\left( \mathrm{B}PGL_3,\mathbf{Z}/3\right) $ being the same as in
part (IV) of the proof of Th. \ref{result1}. So we only need to show that $%
\widetilde{y_8}$ is in the image of $\left( \widetilde{\mathrm{cl}}%
_{PGL_3}\right) _{\left( 3\right) }$. By part (IV) of the proof of Th. \ref
{result1}, $y_8$ is in the image of
\[
j_{\bullet }\circ \left( \mathrm{cl}_{PGL_3}\right) _{\left( 3\right)
}:\left( A_{PGL_3}^4\right) _{\left( 3\right) }\longrightarrow H^8\left(
\mathrm{B}PGL_3,\mathbf{Z}/3\right) ,
\]
and this concludes the proof since $r$ has kernel $BP^{<0}\cdot BP^{*}(%
\mathrm{B}PGL_3)$. \TeXButton{End Proof}{\endproof}
\begin{remark}
We wish to point out that we do not know whether $\ker \widetilde{\mathrm{cl}%
}$ is zero or not. Moreover, since $\mathrm{cl}\left( \chi \right) =0$ and $%
\underline{\mathrm{cl}}$ is injective (Remark \ref{totaro}), we also have $%
\widetilde{\mathrm{cl}}\left( \chi \right) =0$. Therefore, if Totaro's
conjecture was true (i.e $\widetilde{\mathrm{cl}}$ was an isomorphism) we
should have $\chi =0$; but, again, we are not able to prove whether $\chi =0$
or not.
\end{remark}
\section{Appendix. A cohomology-independent proof that $A_{PGL_3}^{*}$ is
not generated by Chern classes}
Here we give an alternative proof of Theorem \ref{result1} which is
independent of Kono-Mimura-Shimada's results on the $\mathbf{Z}/3$%
-cohomology of $\mathrm{B}PGL_3$ and deals only with Chow rings with no
reference to cohomology. However, for the same reason, the following proof
does not yield any direct information on the cycle or refined cycle map.
The notations are those of the previous sections.
\begin{proposition}
\label{RRing}The representation ring of $PGL_3$ is generated by $\left\{
\mathrm{sl}_3,Sym^3E,Sym^3E^{\vee }\right\} $.
\end{proposition}
\TeXButton{Proof}{\proof} The exact sequence
\[
1\rightarrow \mathbf{\mu }_3\longrightarrow SL_3\longrightarrow
PGL_3\rightarrow 1
\]
induces an exact sequence of character groups
\[
0\rightarrow \widehat{T_{PGL_3}}\equiv \widehat{T}\longrightarrow \widehat{%
T_{SL_3}}\stackrel{\pi }{\longrightarrow }\mathbf{Z}/3\rightarrow 0
\]
where $\widehat{T_{SL_3}}=\mathbf{Z}^3/\mathbf{Z}$ ($\mathbf{Z}%
\hookrightarrow \mathbf{Z}^3$ diagonally) and $\pi :\left[
n_1,n_2,n_3\right] \mapsto \left[ n_1+n_2+n_3\right] $. Then
\[
\mathbf{Z}\left[ \widehat{T}\right] \hookrightarrow \mathbf{Z}\left[
\widehat{T_{SL_3}}\right] =\mathbf{Z}\left[ x_1,x_2,x_3\right] \diagup
\left( x_1x_2x_3-1\right)
\]
is the subring generated by monomials $x_1^{n_1}x_2^{n_2}x_3^{n_3}$ with $%
n_1+n_2+n_3\equiv 0$ $\limfunc{mod}3$. Therefore
\[
R\left( PGL_3\right) =\left( R\left( T\right) \right) ^{S_3}=\left( \mathbf{Z%
}\left[ \widehat{T}\right] \right) ^{S_3}\hookrightarrow R\left( SL_3\right)
=\left( \mathbf{Z}\left[ \widehat{T_{SL_3}}\right] \right) ^{S_3}=\mathbf{Z}%
\left[ s_1,s_2\right]
\]
(where $s_i$ is the $i$-th elementary symmetric function on the $x_i$'s) is
the subring generated by $\left\{ s_1^3,s_1s_2,s_2^3\right\} $. Then to
prove the Proposition it is enough to compute $\mathrm{sl}_3,Sym^3E$ and $%
Sym^3E^{\vee }$ in terms of $s_1$ and $s_2$ \emph{in }$R\left( SL_3\right) $.
If $E$ is the standard representation and $\mathbf{1}$ the trivial one
dimensional representation of $SL_3$, we have
\[
E=x_1+x_2+x_3=s_1
\]
\[
E^{\vee }=x_1^{-1}+x_2^{-1}+x_3^{-1}=x_1x_2+x_1x_3+x_2x_3=s_2
\]
\[
\mathrm{sl}_3=E\otimes E^{\vee }-\mathbf{1}=s_1s_2-1;
\]
so
\[
Sym^3E=s_1^3-2s_1s_2+1,\text{ }Sym^3E^{\vee }=s_2^3-2s_1s_2+1
\]
and we conclude. \TeXButton{End Proof}{\endproof}
\begin{corollary}
\label{corRRing}The Chern subring $A_{Ch,PGL_3}^{*}$ of $A_{PGL_3}^{*}$,
generated by Chern classes of representations, is generated by $\left\{
c_i\left( \mathrm{sl}_3\right) ,c_j\left( Sym^3E\right) \right\} _{i,j\geq 0}
$.
\end{corollary}
\begin{theorem}
$\rho $ is not in the Chern subring of $A_{PGL_3}^{*}$.
\end{theorem}
\TeXButton{Proof}{\proof} By Prop. \ref{RRing},
\[
R\left( PGL_3\right) =\mathbf{Z}\left[ \mathrm{sl}_3,Sym^3E,Sym^3E^{\vee
}\right]
\]
and since $\mathrm{sl}_3$ (respectively, $Sym^3E$) is isomorphic to the
regular $A_3\times \mathbf{\mu }_3$-representation minus the trivial one
(respectively, plus the trivial one), we have
\begin{equation}
c_i\left( \mathrm{sl}_3\right) _{\mid A_3\times \mathbf{\mu }_3}=c_j\left(
Sym^3E\right) _{\mid A_3\times \mathbf{\mu }_3}=0,\text{ }i,j=1,2,3,4.
\label{app1}
\end{equation}
Now recall (Section 3) that $\rho $ is a lift to $A_{PGL_3}^{*}$ of
\[
\psi \left( \alpha c_3\left( W\right) \right) \in A_{A_3\ltimes T}^{*}\left(
Diag_{\mathrm{sl}_3}^{*}\right)
\]
where
\[
\psi :A_{A_3\ltimes T}^{*}\longrightarrow A_{A_3\ltimes T}^{*}\left( Diag_{%
\mathrm{sl}_3}^{*}\right)
\]
is the (surjective) pullback. So, the image of $\rho $ under the restriction
\[
A_{PGL_3}^{*}\longrightarrow A_{A_3\ltimes T}^{*}
\]
is of the form $\alpha c_3\left( W\right) +\xi $, for some $\xi \in \ker
\left( \psi \right) $.
Now, let us suppose $\rho $ is in the Chern subring $A_{Ch,PGL_3}^{*}$. By (%
\ref{app1}), we have
\[
\alpha c_3\left( W\right) +\xi \in \ker \left( \varphi :A_{A_3\ltimes
T}^{*}\longrightarrow A_{A_3\times \mathbf{\mu }_3}^{*}\right) .
\]
From the commutative diagram
\[
\begin{tabular}{ccc}
$A_{A_3\ltimes T}^{*}$ & $\stackrel{\varphi }{\longrightarrow }$ & $%
A_{A_3\times \mathbf{\mu }_3}^{*}$ \\
$^\psi \downarrow $ & & $\downarrow ^{\psi ^{\prime }}$ \\
$A_{A_3\ltimes T}^{*}\left( Diag_{\mathrm{sl}_3}^{*}\right) $ & $\stackunder{%
\phi }{\longrightarrow }$ & $A_{A_3\times \mathbf{\mu }_3}^{*}\left( Diag_{%
\mathrm{sl}_3}^{*}\right) $%
\end{tabular}
\]
we get
\[
\psi ^{\prime }\left( \alpha c_3\left( W\right) \right) =0.
\]
Therefore, if we show that $\alpha c_3\left( W\right) $ is not in the kernel
of $\psi ^{\prime }$, we will have proved that $\rho $ cannot be in the
Chern subring of $A_{PGL_3}^{*}$. To do this, let us consider the two
localization sequences\footnote{%
Here $Diag_{\mathrm{sl}_3}$ are the diagonal matrices in $\mathrm{sl}_3$ and
we identify $A_{A_3\times \mathbf{\mu }_3}^{*}\simeq A_{A_3}^{*}\otimes A_{%
\mathbf{\mu }_3}^{*}$ with
\[
\frac{\mathbf{Z}\left[ \alpha \right] }{\left( 3\alpha \right) }\otimes
\frac{\mathbf{Z}\left[ \beta \right] }{\left( 3\beta \right) }.
\]
}:
\begin{equation}
A_{A_3\times \mathbf{\mu }_3}^{*}\stackrel{\cdot (-\alpha ^2)\otimes 1}{%
\longrightarrow }A_{A_3\times \mathbf{\mu }_3}^{*}\left( Diag_{\mathrm{sl}%
_3}\right) \stackrel{p}{\longrightarrow }A_{A_3\times \mathbf{\mu }%
_3}^{*}\left( Diag_{\mathrm{sl}_3}\smallsetminus \left\{ 0\right\} \right)
\rightarrow 0 \label{app2}
\end{equation}
\begin{equation}
A_{\mathbf{\mu }_3}^{*}\simeq A_{A_3\times \mathbf{\mu }_3}^{*}\stackrel{}{%
\left( Z\right) \stackrel{j_{*}}{\longrightarrow }}A_{A_3\times \mathbf{\mu }%
_3}^{*}\left( Diag_{\mathrm{sl}_3}\smallsetminus \left\{ 0\right\} \right)
\stackrel{q}{\longrightarrow }A_{A_3\times \mathbf{\mu }_3}^{*}\left( Diag_{%
\mathrm{sl}_3}^{*}\right) \rightarrow 0 \label{app3}
\end{equation}
where we used that
\[
Z\simeq A_3\times \mathbf{C}^{*},
\]
$A_3\ltimes T$-equivariantly. Since (Section 3),
\[
W\simeq \mathbf{C}_{\chi ,\mathbf{\mu }_3}\boxtimes \mathbf{C}_{perm,A_3}^3
\]
as $A_3\times \mathbf{\mu }_3$-representations (where $\mathbf{C}_{\chi ,%
\mathbf{\mu }_3}$ is the $\mathbf{\mu }_3$-representation of character $\chi
=\exp (i2\pi /3)$ and $\mathbf{C}_{perm,A_3}^3$ is the $A_3$-permutation
representation), its Chern roots are
\[
\left\{ \beta +\alpha ,\beta -\alpha ,\beta \right\}
\]
and then
\begin{equation}
\alpha c_3(W)_{\mid A_3\times \mathbf{\mu }_3}=(\beta ^2-\alpha ^2)\alpha
\beta . \label{app4}
\end{equation}
By (\ref{app4}) and (\ref{app2}), it is enough to prove that $j_{*}=0$.
Let us consider the pullback $E$ of $\mathbf{C}_{\chi ,\mathbf{\mu }_3}$ to $%
X=Diag_{sl_3}\smallsetminus \left\{ 0\right\} $ as an $A_3\times \mathbf{\mu
}_3$-equivariant vector bundle, with $A_3$ acting trivially on $\mathbf{C}%
_{\chi ,\mathbf{\mu }_3}$ and $A_3\times \mathbf{\mu }_3$ acting as usual on
$X$ (i.e. $\mathbf{\mu }_3$ acting trivially and $A_3$ by permutations). We
have
\[
j^{*}\left( c_1\left( E\right) \right) \equiv c_1(E)_{\mid \mu _3}=\beta .
\]
But we also have $j_{*}\left( 1\right) =0,$ since
\[
Z=D^{-1}\left( \left\{ 0\right\} \right)
\]
where
\begin{eqnarray}
D &:&X\rightarrow \mathbf{A}^1 \nonumber \\
\left( \lambda _1,\lambda _2,\lambda _3\right) &\longmapsto &\left( \lambda
_1-\lambda _2\right) \left( \lambda _1-\lambda _3\right) \left( \lambda
_2-\lambda _3\right)
\end{eqnarray}
is the square root of the discriminant (which is $A_3\times \mathbf{\mu }_3$%
-equivariant). So $j_{*}=0$ and we conclude. \TeXButton{End Proof}{\endproof}%
\
|
\section{Introduction}
At a near fundamental level, liquid metals are complex binary fluids
consisting of ions in a sea of conduction electrons, their physical
properties linked to the three corresponding correlation functions:
$S_{II}(k), S_{eI}(k)$, and $S_{ee}(k)$\cite{Ashc78}. We present a simple and
evidently accurate analytic scheme to calculate electron-ion
structure factors ($S_{eI}(k)$) by combining a hard-sphere
approximation for the ionic structure with a simple linear
response theory for the electrons. These structure factors are now in
principle accessible experimentally through recent advances in both
neutron and x-ray scattering techniques. Another route to effective
electron-ion interactions therefore opens, but now through the fluid
state.
We also address the evident success of the linear approximation
by studying a related problem,
the density, $\rho^{ind}({\bf \vec k})$, of an initially uniform
electron gas induced by an embedded pseudo-potential, $v^{ps}({\bf \vec
k})$. By comparing linear and second order response to full
(Kohn-Sham\cite{Kohn65}) non-linear response, we show that even though
$v^{ps}({\bf \vec k})$ is not necessarily a small perturbation,
the consequent response series converges term by term.
The non-linear terms are significantly reduced by an
interference between atomic and electronic length scales for most
metals, the main exception to this being hydrogen.
\section{Discussion}
\subsection{Electron-ion correlation functions}
The electron-ion-structure factor can be written as\cite{Ashc78}:
\begin{equation}\label{eq5}
S_{eI}(k) = \frac{1}{\sqrt{N_e N_I}}
< \hat{\rho}_e^{(1)}({\bf \vec k}) \hat{\rho}_I^{(1)}({\bf \vec k}) >
= \frac{n(k)}{\sqrt{Z}} S_{II}(k),
\end{equation}
where $\hat{\rho}_j^{(1)}(k)$ is the Fourier transform
of the one-particle density operator of component $j$,
$S_{II}(k)$ is the ion-ion structure factor and $n(k)$
is identified as the
pseudo-electron density, or pseudo-atom (of valence Z). Thus electron-ion
correlations can be described by convolving
the pseudo-atom with the ionic correlations. The ionic correlations
are themselves well described by a Percus-Yevick hard-sphere structure
factor\cite{Ashc66b}, while the pseudo-atom is described by a standard
linear response formulation:
\begin{equation}\label{eq5a}
n(k) = \chi_1(k) v^{ps}(k)
\end{equation}
where $\chi_1(k)$ is the well known linear response function; to approximate
$\chi_1(k)$ we use a Local Density Approximation (LDA)
local field factor\cite{Hafn87}.
The electron-ion interaction is modeled by a simple local
one-parameter empty-core pseudopotential\cite{Ashc66} i.e.: $v^{ps}(k) =
-(4 \pi e^2/k^2) cos(kR_c)$, where $R_c$ is the core radius
and the pseudo-potential goes through zero at $k_0 = \pi/2 R_c$. (We note that
at this linear level, the effects of ionic averaging on the pseudo-atom
are ignored\cite{thesis}).
Using the approximations above in~(\ref{eq5}), we compare our
approach in Fig. 1 to the {\em full ab-initio} Car-Parrinello\cite{Car85} calculations
of de Wijs {\em et al}\cite{deWi95}. The correspondence is striking,
especially when we note that the parameters $\eta$ and $R_c$ are {\em
a priori} set by other physical properties (no fitting is necessary).
Besides a semi-quantitative description of electron-ion structure
factors, this linear response theory now provides an important
qualitative insight into the form of the electron-ion structure
factors\cite{Loui97b}. The pseudo-atom density, $n(k)$, is typically
largest for smaller $k$ and thereafter
rapidly declines for larger $k$, while the
near classical ion-ion structure factor, $S_{II}(k)$, follows an inverse
behavior; it is small for small $k$. Together with the product
form~(\ref{eq5}) this implies that the shape of the electron-ion
structure factor, $S_{eI}(k)$, is determined primarily by the the
position of the {\em zero-crossing}, $\bar{k}_0$, of $n(k)$ with respect
to the {\em first maximum}, $k_p$, of $S_{II}(k)$. If $\bar{k}_0 <
k_p$, then $S_{II}(k)$ selects (or filters) the negative part of
$n(k)$ and $S_{eI}(k)$ takes a form similar to that of Mg (Fig. 1
(a)). Conversely, if $\bar{k}_0 > k_p$, then the ion-ion structure
factor selects (or filters) the positive part of $n(k)$, and again,
$S_{eI}(k)$ takes a form similar to that of Bi (Fig. 1
~(\ref{figfullBi_Mg} (c)). Since $\chi_1(k)$ is positive definite,
the zero-crossing in linear response occurs at $k_0$. The large slope
of $n(k)$ near the zero-crossing then implies that non-linear
corrections {\em must} have a small effect on the location of the
zero-crossing, and this, together with the expected accuracy of linear
response, implies that $\bar{k}_0 \sim k_0$. For most metals,
$k_0$ is just a little less than $2k_F$, and the latter's ratio to
$k_p$ is well known: for low valence $(Z \leq 2)$, $2k_F < k_p$;
for high valence $(Z\geq 3)$: $2k_F > k_p$\cite{Zima72}. This
accounts in a straightforward way for the two separate forms found by
deWijs {\em et al}\cite{deWi95}: For Mg, $\bar{k}_0 < k_p$ $(Z=2)$,
which belongs to the {\bf low valence class} of electron-ion structure
factors. For Bi, $\bar{k}_0 > k_p$ $(Z=5)$ and we may refer to this
as the {\bf high valence class} of electron-ion structure
factors\cite{deWijssei}. Generally ions of valence $Z \leq 2$ belong
to the low valence class while ions with valence $Z > 3 $ belong to
the high valence class. Ions with valence $Z=3$ typically belong to
the high valence class also, although they may be characterized by a
crossover form\cite{thesis}. The analytical approach above can easily
be extended by using the modern theory of classical liquids
to obtain
improved ion-ion structure factors\cite{Cusa76}, but to include second
order contributions to the pseudo-atom $n(k)$ necessitates not only
second order electron response, but also contributions from ion-ion
triplet structure. The latter can be carried out with concepts from
the theory of classical liquids\cite{thesis}.
\subsection{Proposed Experiments}
The principal features of electron-ion structure factors can be
measured by exploiting the differences between the
x-ray scattering structure factor, $S_{II}^X(k)$, determined with
a free-atom form factor, $f_A(k)$, and the structure factor, $S_{II}^N(k)$,
determined by neutron scattering\cite{Egel74,Chih87}. As emphasized
by Chihara\cite{Chih87} the x-ray
structure factor for liquid metals will equal $S_{II}^N(k)$ only when
determined with an
ionic form factor augmented by a pseudo-atom form factor; i.e.
$f_I(k) + n(k)$, so that:
\begin{equation}\label{eq10}
\frac{S_{II}^X(k)}{S_{II}^N(k)} = \frac{|f_I(k) + n(k)|^2}{|f_A(k)|^2}.
\end{equation}
This effect is
clearly expected to be largest for metals with larger ratios of
valence to core electrons. Thus we predict a small effect for metals
with smaller valence to core ratio such as Na or K,
a 2\% difference at the 1st peak of the structure for Li (ratio$=1:2$)
or Al (ratio$=3:10$), but by far the largest effect for Be (ratio $=1:1$)
where the difference at the
principal peak of the structure factor could be as high as 7\%, well
within experimental range. Another interesting candidate would be
metallic Si (ratio $=4:10$) since covalent effects still make
themselves felt in the liquid state suggesting that experiments
could reveal effects beyond linear response.
To date the experimental electron-ion structure factors and related
pseudo-atoms show
considerably more structure than indicated by theoretical
predictions\cite{Take85}.
Significant experimental challenges are faced
in the
accuracy resulting from subtraction of two sets of data obtained by quite
different means, each with important (but different) systematic corrections;
however the present approach suggests that the current differences between
x-ray and neutron scattering should be reexamined (see also\cite{Chih87}).
Using a pseudo-atom instead of the full free atom as
a form factor can assist in comparing neutron and x-ray measurements
and help unravel various systematic corrections applied.
The advent of high precision x-ray and neutron sources currently
coming on-line suggests that these proposed effects can be systematically
explored.
\subsection{Non-linear response of an atom in an electron gas}
The evident (and long-standing) success of the linear response approximation
for electron response\cite{Hafn87}, here demonstrated for
electron-ion structure factors, calls for further investigation. The
accuracy of linear response in a crystalline solid is commonly
attributed to the fact that the structure dependent reciprocal lattice
vectors are typically near the pseudo-potential zero-crossing, $k_0$,
with the associated inference that the net scattering is smaller than one would
naively expect\cite{Hafn87}. For
liquids or other disordered systems such arguments are less appropriate. To
examine the strength of linear response in the absence of ionic structure,
we consider a simpler problem, namely the
response of the interacting electron gas to a single ion, where the
electron-ion interaction is modeled by a simple local one-parameter
empty-core pseudopotential\cite{Ashc66}. We will compare two routes
to the induced density, $\rho^{ind}(k)$. The first follows from
solving the Kohn-Sham equations\cite{Kohn65} exactly (for the given pseudopotential)
within the local density
approximation (LDA), the second from the standard expansion
of the
response in powers of the perturbing (pseudo)potential, i.e.;
\begin{equation}\label{eq1}
\rho^{ind}(k) = \chi_1(k) v^{ps}(k) +
\sum_{\vec{k}_1} \sum_{\vec{k}_2}
\chi_2(k,k_1,k_2)v^{ps}(k_1)v^{ps}(k_2) + \cdots \, .
\end{equation}
Here the response functions, $\chi_n(k_1...)$, are properties of the
{\em homogeneous} interacting electron gas. In Fig. 2 we compare
the explicit second order response with an LDA local field
factor\cite{thesis,Lloy68} to the full non-linear LDA response.
Clearly the
non-linear response is well characterized by the 2nd order term,
implying that the success of linear response is not due to
cancellation between higher order terms of opposite sign but instead
that each successive term is individually small compared to the previous
term in the expansion; {\em the response series
converges very rapidly, term by term}.
The non-linear response is largest for atomic parameter $R_c =0$ (hydrogen), and
decreases with a larger atomic-parameter, $R_c$, as might be physically
anticipated. However as $R_c$ increases from zero, a noticeable secondary
minimum occurs when the inverse atomic length, $k_0$, is equal to $2k_F$. For
the cases plotted in Fig. 2, the maximum in second order response at $k_0/2k_F
= 1$ (or $R_c/r_s = 0.41$) is reduced by an entire {\em order of magnitude} when
compared with the maximum in second order response calculated for hydrogen
($R_c/r_s = 0$), and is typically equal to the value at $3 R_c/r_s$. The
physics behind this minimum is attributed to the following; the second order
response function, $\chi_2(k,k_1,k_2)$, peaks when the summed arguments
in~(\ref{eq1}) are close to $2k_F$\cite{thesis}. If the pseudo-potential
zero-crossing, $k_0$, is near the response peaks at $2k_F$, a maximal
cancellation or {\em maximal destructive interference of the atomic and
electronic length scales} occurs, leading to a minimum in second order response.
The ratio of the atomic and electronic length scales is set primarily by the
volume energy terms in the total ground state energy, and is almost independent
of structure\cite{Hafn87}; $k_0/2k_F$ lies between $0.75$ and $1$ for most
metals, and is therefore very close to the secondary minimum in the non-linear
response.
As noted, the effect we discuss originates from an {\em interference} between
intrinsic {\em atomic} and {\em electronic} length scales, but it also
complements the argument given for crystalline solids alone, which stems from
the {\em confluence } of an {\em atomic} and a {\em structural} length scale.
The clear exception to these interference effects is again the singular case of
a point-charge $(v^{ps}(k) \sim 4 \pi e^2/k^2)$, i.e. the case of hydrogen,
which has no well-defined core-length scale, $k_0$, no oscillations in the
potential and thus no interference effect in the higher order terms. In
contrast to other systems, non-linear response terms are large {term by term}.
In fact, the response series may not even formally converge and care must be
taken when applying concepts derived from linear-response theory to hydrogen (it
is not a simple material).
Finally we note that the second
order response contribution is of the same order as
the difference between first order response with or without local field
corrections. In addition, the combined effects of exchange and correlation
partially cancel between first and second order, implying that {\em neglect of
higher order response results in an over-estimation of the role of exchange and
correlation}, which, in turn, has important implications for the widespread
application of linear response theory in the derivation of effective ion-ion
potentials in (simple) metals.
\section{Conclusions}
The electron-ion structure factors of liquid simple-metals
are well described by a
simple linear response theory augmented by linear response for the electrons.
This approach suggests two main classes of electron-ion correlation functions,
one for high and one for low valence metals. Experimental advances in x-ray
and neutron-scattering may be able to provide measurements of these
electron-ion correlation functions, with liquid Be being the most promising
candidate. A route to information on fundamental electron-ion interactions
therefore becomes available through the fluid state. Finally, the well
documented success of the linear response approximation for electrons stems in
part from an interference effect between atomic and electronic
length scales.
This work was supported by the NSF through the Cornell Center for
Materials Research under Grant No. DMR96-32275. We
especially thank Professor Karsten Jacobsen for making an
LDA Kohn-Sham program
available to us, and Dr David Muller for helpful suggestions.
|
\section{Introduction}
The study of strangeness production in relativistic heavy-ion
collisions has been of continuing interest as strangeness
is predicted to be enhanced by the formation of a quark gluon plasma
(QGP)\cite{Chang_QGP}. At the same time, many particle properties
such as effective
mass, production threshold, and absorption cross section may be sensitive
to a high baryon density.
Their study might reveal the influence of a many-body mean-field
potential and provide a signal of chiral symmetry restoration.
Experiment E917 at the AGS measured Au+Au collisions at beam kinetic
energies of 6, 8 and 10.8 AGeV in the winter of 1996/97. The Henry Higgins
spectrometer, used previously in experiments E802, E859 and
E866 \cite{Chang_e866kaon}, and an
upgraded data acquisition system enabled the experiment
to take 280$\times10^6$
kaon-pair/$\overline{p}$ triggered events. The quality of this data set
allows for a detailed study of short lived vector mesons, baryons,
anti-baryons and the systematics of two particle correlations of pion,
kaon and proton pairs. E917 is unique among the AGS experiments in its
ability to measure a wide variety of strangeness-carrying particles
including $K^+$, $K^-$, $\Lambda$, $\overline{\Lambda}$ and $\phi$-mesons. The study of
the excitation function of kaon production may help identify a possible
phase transition, and the study of $\phi$-mesons may
provide a direct probe of any in-medium effect. More details on the
experimental setup and trigger condition can be found in
Ref.~\cite{Chang_birger,Chang_jamie}.
This article presents a systematic study
of the spectra and yield of $K^{+}\:$ and $K^{-}\:$ in Au+Au collisions at beam
energies of 6 and 8 AGeV combined with published E866
data~\cite{Chang_e866kaon} at 10.8 AGeV. A previously reported
discrepancy in the
measured $\overline{p}$ yields between AGS experiments E878 and E864 has been
hypothesized to arise from an abundance of $\overline{\Lambda}$ production and
different acceptances for the $\overline{p}$ daughter from $\overline{\Lambda}$ decay
in the two experiments.
Experiment E917 is able to make the first direct measurement of
$\overline{\Lambda}$ yields, for which preliminary results are presented.
\section[Kaon Production]{Measurement of Kaons}
There is significant theoretical interest in the study of kaon
properties in dense nuclear matter. Qualitatively, the models
suggest that $K^{+}\:$ mesons experience a weak repulsive potential
inside the nuclear medium resulting in a slight increase in the
effective mass with baryon density, whereas
a strong attractive potential for $K^{-}\:$ mesons leads to a
significantly reduced effective mass in the high baryon density
environment \cite{Chang_Weise}. Because
the $\Lambda K^{+}\:$ production channel is expected to be
larger than $K^-K^{+}\:$ pair production at AGS energies,
this scenario results in
a larger $K^{-}/K^{+}\:$ yield ratio for central events near mid-rapidity, where
high baryon density is expected and the $K^-K^{+}\:$ channel is
enhanced.
The rapidity distributions of kaons were obtained from exponential fits to
the transverse mass, $m_t$, spectra at each rapidity bin. From these
fits we obtain the integrated production probability, $dN/dy$, per unit of
rapidity and the inverse slope, $T_{inv}$, of these spectra. We emphasize
that $T_{inv}$ should not be interpreted as the {\it
temperature} of the emitting source as it is well known that collective
effects, such as radial expansion, can mimic high source temperatures.
The rapidity distributions, $dN/dy$, for $K^{+}\:$ and $K^{-}\:$ emission
for 0-5\% central collisions are shown in Fig.~\ref{Chang_fig1}
for beam energies of 6, 8 and 10.8 ~AGeV .
The rapidity distributions are observed to be peaked
at mid-rapidity and the yields increase with beam energy without
substantial change in the shape of the rapidity distributions.
We also find that the rapidity distributions
are essentially independent of centrality.
\begin{figure}[ht]
\centerline{
\epsfig{file=chang_fig1.eps,angle=0,width=\textwidth}}
\caption{Rapidity distributions for $K^{+}\:$ (panel a) and $K^{-}\:$ (panel b)
for 0-5\% central events are shown for beam energies of 6, 8
(E917 preliminary) and 10.8 AGeV (E866 \cite{Chang_e866kaon}).}
\label{Chang_fig1}
\end{figure}
Since the effects of the nuclear medium have opposite sign for $K^{+}\:$ and $K^{-}\:$,
the ratio of $K^{-}/K^{+}\:$ production might be a very sensitive probe for studying
such effects. It was studied as a function of global parameters for the
collision, such as centrality and rapidity. In
Fig.~\ref{Chang_fig3}(a), the $K^{-}/K^{+}\:$ ratio is shown as a function of
rapidity for the 5\% most central events at 6, 8, and
10.8 \cite{Chang_e866kaon} AGeV.
We observe that the ratio increases with beam energy
over the rapidity range of this study
and that the rapidity distribution for $K^{-}\:$
is narrower than for $K^{+}\:$, an observation that has also been made in studies of
Ni+Ni collisions at SIS energies~\cite{Chang_fopi_ni}.
The fact that the production of $K^{-}\:$ relative to
$K^{+}\:$ is more abundant around mid-rapidity might be expected, because:
\begin{itemize}
\item the available energy for producing particles is peaked around
mid-rapidity.
\item the baryon density is largest around mid-rapidity and the in-medium
effect enhances the production of $K^{-}\:$ relative to that of $K^{+}\:$.
\end{itemize}
It is, however, difficult to disentangle the relative importance of these two
effects~\cite{Chang_gqli}.
The measurements of the $K^{-}/K^{+}\:$-ratio at mid-rapidity
is shown in
Fig.~\ref{Chang_fig3}(b) as a function of center-of-mass
energy from SIS through AGS to SPS energies.
The observed increase in the ratio with beam energy may be expected on the
basis of the higher production threshold for $K^-$. This makes the production
cross section of $K^{-}\:$ increase faster than that of $K^{+}\:$ above the
production threshold~\cite{Chang_gqli}. At SPS energy, the increase in the
ratio is not as steep as that in the lower energy. This is probably
caused by a near saturation in the population of the available phase space for
both $K^{+}\:$ and $K^-$.
\begin{figure}[ht]
\centerline{
\epsfig{file=chang_fig2.eps,width=\textwidth}}
\caption{ The rapidity distribution of $K^{-}/K^{+}\:$ ratio for the 0-5\%
central events at various beam energies (panel a) and the $K^{-}/K^{+}\:$ ratio at
mid-rapidity at SIS, AGS and SPS~\cite{Chang_fopi_ni, Chang_qm97} (panel b).}
\label{Chang_fig3}
\end{figure}
We have also studied the centrality dependence
of $K^{-}/K^{+}\:$ over a wide range of rapidities.
The rapidity distribution of this ratio
exhibits a very weak dependence on centrality at all three energies,
similar to
the observations at SIS energies~\cite{Chang_fopi_ni, Chang_kaos_ni}.
This weak
dependence seemingly contradicts the naive expectation based on the
in-medium effect. Thus, one might expect that the $K^{-}/K^{+}\:$ ratio
at mid-rapidity should increase
significantly towards central collisions, and that this increase
should be
most pronounced in the low-energy region close to the production
threshold where the effect of the reduction (increase) of $K^{-}\:$ ($K^{+}\:$)
mass from the in-medium effect is expected to be strongest. Li and Brown
have proposed~\cite{Chang_gqli} that a suppression
of $K^{-}\:$ production through the hyperon-feeding channel compensates for
the increase in the $K^{-}/K^{+}\:$-ratio from the in-medium effect in the central
collisions. It is, however, difficult to verify this hypothesis
experimentally.
The inverse slope parameter, $T_{inv}$, derived from the fits to the
$m_t$-spectra is found to peak at mid-rapidity for both $K^{+}\:$ and
$K^{-}\:$. There is also a slight increase in $T_{inv}$ with beam energy, but the
rapidity dependence is virtually unchanged.
The difference in $T_{inv}$ for $K^{+}\:$ and $K^{-}\:$ transverse mass spectra
is found to be small, although the value of $T_{inv}$ for $K^{+}\:$ is
about 50 MeV larger than that of
$K^{-}\:$ at the 6 AGeV beam energy.
\section{Measurements of $\overline{\Lambda}$ and $\overline{p}$}
AGS experiment E859 has measured a large $\overline{\Lambda}/\overline{p}$ ratio of $2.9 \pm 0.9
\pm 0.5$ for Si+Au at 13.7 AGeV~\cite{Chang_e859}.
This ratio is unexpectedly large
relative to thermal model calculations or with reference to the results for NN
collisions at AGS energies~\cite{Chang_welke}.
In addition, experiments E864 and E878 at the AGS have
measured $\overline{p}$ production in the
mid-rapidity region and zero $p_t$ for Au+Pb at 10.6 AGeV~\cite{Chang_e864}
and Au+Au at 10.8 AGeV~\cite{Chang_e878}
respectively. There is a significant discrepancy between the two experiments
in the reported
anti-proton production probability for central
events (about a factor of 3.5), but a good agreement for the most
peripheral events. Since these two experiments have different
acceptances for detecting the $\overline{p}$ from $\overline{\Lambda}$ and
$\overline{\Sigma}$ decay, a large production of $\overline{\Lambda}$
might reconcile the results for the two experiments. If this discrepancy
is attributed entirely to this effect,
a $\overline{\Lambda}/\overline{p}$ ratio of 3.5 (most probable value) or larger
than 2.3 (98\% confidence level) is required.
\begin{figure}[ht]
\centerline{
\epsfig{file=chang_fig3,height=8cm}}
\caption{The invariant mass distribution of $\overline{\Lambda}$ reconstructed from
the pair of $\overline{p}$ and $\pi^{+}$. The line is the fitted background
from mixed events.}
\label{Chang_fig6}
\end{figure}
\begin{table}[hbt]
\caption[The various quantities]{The measurement of $\overline{p}$ and $\overline{\Lambda}$
by E917 in the rapidity interval $y=1.0-1.4$.}
\begin{tabular*}{\textwidth}{@{\extracolsep{\fill}}lcccc}
\savehline \savehline
& \multicolumn{2}{c}{Minimum bias} &
\multicolumn{2}{c}{Central events 0-23\%} \\ \hline
Particle & $dN/dy$ & $T_{inv}$ & $dN/dy$ & $T_{inv}$ \\
& $(\times 10^{-3})$ & (MeV) & $(\times 10^{-3}$) & (MeV) \\
\hline
\vspace{5pt}
$\overline{\Lambda}^a$ &$4.3^{+1.8}_{-1.2}$ & $243^{+112}_{-59}$ & $12.9^{+5.5}_{-3.7}$& $243^{+110}_{-60}$\\
\vspace{5pt}
$\overline{p}_{measured}$ & $7.31 \pm 0.17$ & $179 \pm 8$ & $15.0 \pm 0.6$ & $196 \pm 11$ \\
\vspace{5pt}
$\overline{p}_{direct}$ & $4.6^{+0.7}_{-1.2}$ & & $6.8^{+2.3}_{-3.6}$ &\\
\hline
\vspace{5pt}
Ratio $\overline{\Lambda}/\overline{p}_{direct}$& \multicolumn{2}{c}{$0.9^{+0.9}_{-0.3}$} &
\multicolumn{2}{c}{$1.9^{+3.8}_{-0.9}$} \\
\savehline
\end{tabular*}
\begin{tablenotes}
$^a$ dN($\overline{\Lambda}$)/dy = dN($\overline{\Lambda} \rightarrow \overline{p} \pi^+$)/dy / 0.64. \\
\end{tablenotes}
\label{Chang_tab1}
\end{table}
Experiment E917 measured $\overline{p}$ in the rapidity range $1.0<y<1.4$ and
$\overline{\Lambda}$ were reconstructed from $\overline{p} \pi^{+}$ pairs.
The signal of $\overline{\Lambda}$ is clearly
seen in the invariant mass distribution shown in the
Fig.~\ref{Chang_fig6}. The transverse mass spectra of $\overline{p}$ and
$\overline{\Lambda}$ in the rapidity range $1.0<y<1.4$ for the central 0-23\%
events are shown in Fig.~\ref{Chang_fig7}. The efficiency of detecting
$\overline{p}$ from $\overline{\Lambda}$ decay is close to unity in our experiment. Assuming
that the decay of $\overline{\Lambda}$ is the only source of hyperon feed-down into
$\overline{p}$, the yield of $\overline{p}$ directly produced is
$dN(\overline{p}_{\rm direct})/dy = dN(\overline{p}_{\rm measured})/dy-0.64 dN(\overline{\Lambda})/dy$,
thereby correcting
for the 64\% branching ratio of the $\overline{\Lambda}$ decay into the $\overline{p}\pi^+$ channel.
The rapidity yield, dN/dy, for $1.0<y<1.4$ and the inverse
slope parameter, $T_{inv}$, obtained from a fit to the $m_t$-spectra
with an exponential function, are listed in Table~\ref{Chang_tab1}.
Details on this analysis are available in Ref.~\cite{Chang_george}.
\begin{figure}[hb]
\centerline{
\epsfig{file=chang_fig4.eps,height=8cm}}
\caption{Transverse mass ($m_t$) spectra of $\overline{p}$ (open circles) and
$\overline{\Lambda}$ (solid circles) for the 0-23\% central events.}
\label{Chang_fig7}
\end{figure}
The E917 measurement of the $\overline{\Lambda}/\overline{p}$-ratio is greater than unity for
the 0-23\% central collisions and consistent with the E859 measurement in
Si+Au system. The ratio derived from the difference between the E864 and
E878 $\overline{p}$ measurements lies within the upper bound of
E917 results. It should be noted, however, that there exist several
differences in the experimental measurements presented here and those
of E864, E878, and E859, as listed in Table~\ref{Chang_tab2}. Most
important are the ranges in $m_t$, rapidity, and centrality measured
by the different experiments. More data will be analyzed in the
future to compare the results from the other experiments under similar
centrality and rapidity cuts.
\begin{table}[th]
\caption[The difference]{The difference in experimental conditions for
measuring $\overline{\Lambda}/\overline{p}$-ratio.}
\begin{tabular*}{\textwidth}{@{\extracolsep{\fill}}cccccc}
\savehline
\savehline
Exp. & Collision & $E_{beam}$ & Centrality & Rapidity & $m_t-m_0$ \cr
& & (GeV/nucleon)& (\%) & & (MeV/c$^2$) \\
\savehline
E859 & Si+Au & 13.7 & 0-15 & $1.0-1.4 $ & $> 250$ \cr
E864 & Au+Pb & 10.6 & 0-10 & $1.6-2.0 $ & $= 0$ \cr
E878 & Au+Au & 10.8 & 0-10 & $1.4-2.4 $ & $= 0$ \cr
E917 & Au+Au & 10.8 & 0-23 & $1.0-1.4 $ & $> 250$ \cr
\savehline
\end{tabular*}
\label{Chang_tab2}
\end{table}
\section{Summary and Outlook}
A complete measurement of kaon production at 6 and 8 AGeV has been
presented. No dramatic change is evident in the excitation function of
kaons from 6 to 10.8 AGeV. A straightforward expectation based on the
scenario of many-body in-medium effect on kaons cannot explain the
observation of a weak centrality dependence of $K^{-}/K^{+}\:$ ratios with changing
beam energy.
The $\overline{\Lambda}/\overline{p}$ ratio was measured to be greater than unity for 0-23\%
central collisions. More data need to be analyzed to enable a
detailed comparison with the other results.
The results presented in this talk are all very preliminary in nature.
For this reason, we have not presented any comparison with, or analysis
in terms of, theoretical models.
These will be presented in future publications.
\begin{acknowledgments}
This work was supported by the Department of Energy (USA), the National
Science Foundation (USA), and KOSEF (Korea).
\end{acknowledgments}
\begin{chapthebibliography}{1}
\bibitem{Chang_QGP}
Koch, P., M\"{u}ller, B., and Rafelski, J. (1986) Phys. Rep. {\bf 142}, 167.
\bibitem{Chang_e866kaon}
Ahle, L. {\it et al.}, (1998) Phys.\ Rev.\ {\bf C58}, 3523.
\bibitem{Chang_birger}
Back, B. (1999) these proceedings.
\bibitem{Chang_jamie}
Dunlop, J. C. (1999) Ph.D. thesis, MIT.
\bibitem{Chang_Weise}
Weise, W. (1996) Nucl. Phys. {\bf A610}, 35c.
\bibitem{Chang_fopi_ni}
Hong, B. (1998) in {\it Proceedings of APCTP Workshop on
Astro-Hadron Physics}, edited by G. E. Brown, World Scientific,
Singapore.
\bibitem{Chang_gqli}
Li, G. Q. and Brown, G. E. (1998) Phys.\ Rev.\ {\bf C58}, 1698.
\bibitem{Chang_qm97}
Bearden, I. G. {\it et al.} (1998) Nucl. Phys. {\bf A638}, 419.
\bibitem{Chang_kaos_ni}
Barth, R. {\it et al.} (1997) Phys.\ Rev.\ Lett. {\bf 78}, 4027.
\bibitem{Chang_e859}
Stephans, G. S. and Wu, Y. (1997) J. Phys. G {\bf G23}, 1895.
\bibitem{Chang_welke}
Wang, G. J. {\it et al.} (1998) Los Alamos Preprint Archive nucl-th/9807036
and nucl-th/9806006.
\bibitem{Chang_e864}
Armstrong, T. A. {\it et al.}, (1997) Phys.\ Rev.\ Lett. {\bf
79}, 3351; Los Alamos Preprint Archive nucl-ex/9811002.
\bibitem{Chang_e878}
Beavis, D. {\it et al.} (1995) Phys.\ Rev.\ Lett. {\bf 75}, 3633;
(1997) Phys.\ Rev.\ {\bf C56}, 1521.
\bibitem{Chang_george}
Heintzelman, G. (1999) Ph.D. thesis, MIT.
\end{chapthebibliography}
\end{document}
|
\section{Introduction}
Angular momentum losses to gravitational radiation are expected to
lead to the coalescence of binary systems containing black holes,
and/or neutron stars (when the initial binary separation is small
enough for the decay to take place in less than the Hubble time). This
type of evolution has been suggested in a variety of contexts as
possibly giving rise to observable events, such as gamma---ray bursts
(GRBs) and bursts of gravitational waves (see
e.g. Thorne~1995). Additionally, it could help explain the observed
abundances of heavy elements in our galaxy~(Lattimer \& Schramm~1974;
1976) if the star is tidally disrupted in the encounter (see
Wheeler~1971). Study of such events could also provide constraints on
the equation of state at supra--nuclear densities.
After the recent measurement of redshifts to their
afterglows~\cite{metzger,kulkarni,djorgovski}, it is now generally
believed that GRBs originate at cosmological distances. The calculated
event rates~\cite{latt,nara,tutukov,lipunov,portyun,tomek} for merging
compact binaries are compatible with the observed frequency of GRBs
(on the order of one per day). The preferred model for the production
of a GRB invokes a relativistic fireball from a compact `central
engine' that would produce the observable $\gamma$--rays through
internal shocks~(M\'{e}sz\'{a}ros \& Rees~1992; 1993). This model
requires the presence of a relatively baryon--free line of sight from
the central engine to the observer along which the fireball can expand
at ultrarelativistic speeds. Additionally, the short--timescale
variations seen in many bursts (often in the millisecond range)
probably arise within the central engine~\cite{sari}.
The coalescence of binary neutron star systems or black hole--neutron
star binaries was suggested as a mechanism capable of powering the
gamma--ray bursts, either during the binary merger itself or through
the formation of a dense accretion disk which could survive long
enough to accomodate the variable timescales of GRBs~(Paczy\'{n}ski
1986; Goodman 1986; Goodman, Dar \& Nussinov 1987; Eichler et
al. 1988; Paczy\'{n}ski 1991; M\'{e}sz\'{a}ros \& Rees 1992; Woosley
1993; Jaroszy\'{n}ski 1993; 1996; Witt et al. 1994; Wilson, Mathews \&
Marronetti 1996; Ruffert, Janka \& Sch\"{a}fer 1996; Lee \&
Klu\'{z}niak 1997; Ruffert, Janka, Takahashi \& Sh\"{a}fer 1997; Katz
1997; Klu\'{z}niak \& Lee 1998; Ruffert \& Janka 1998; Popham, Woosley
\& Fryer 1998; McFadden \& Woosley 1998; Ruffert \& Janka 1999). The
enormous amount of gravitational energy that would be liberated in
such an event could account for the energetics of the observed GRBs
and neutrino--antineutrino annihilation may power the necessary
relativistic fireball.
In previous work~(Lee \& Klu\'{z}niak 1995; 1998 (hereafter Paper~I);
Klu\'{z}niak \& Lee 1998), we have studied the coalescence of a
neutron star with a stellar--mass black hole for a stiff ($\Gamma=3$)
polytropic equation of state and a range of mass ratios. We found that
the neutron star was not entirely disrupted, but rather remained in
orbit (with a greatly reduced mass) about the black hole after a quick
episode of mass transfer. Thus the duration of the coalescence process
would be extended from a few milliseconds to possibly several tens of
milliseconds. The observed outcome seemed favorable for the production
of a GRB since in every case we found a baryon--free axis in the
system, along the axis of rotation.
In the present paper, we investigate the coalesence of a black
hole--neutron star binary for a soft equation of state (with an
adiabatic index $\Gamma=5/3$) and a range of mass ratios. Our initial
conditions are as in Paper~I in that they correspond to tidally locked
binaries. Complete tidal locking is not realistically
expected~\cite{bildsten}, but it can be considered as an extreme case
of angular momentum distribution in the system. In the future we will
explore configurations with varying degrees of tidal locking.
As before, the questions motivating our study are: Is the neutron star
tidally disrupted by the black hole and does an accretion torus form
around the black hole? If so, how long--lived is it? Is the baryon
contamination low enough to allow the formation of a relativistic
fireball? Is any significant amount of mass dynamically ejected from
the system? What is the gravitational radiation signal like, and how
does it depend on the equation of state and the initial mass ratio?
In section~\ref{method} we present the method we have used to carry
out our simulations. This is followed by a presentation of our results
in section~\ref{results} and a discussion in section~\ref{discussion}.
\section{Numerical method}\label{method}
For the simulations presented in this paper, we have used the method
known as Smooth Particle Hydrodynamics (SPH). Our code is
three--dimensional and essentially Newtonian. This method has been
described often, we refer the reader to
Monaghan~\shortcite{monaghan92} for a review of the principles of SPH,
and to Paper~I and Lee~\shortcite{phd} for a detailed description of
our own code, including the tree structure used to compute the
gravitational field.
We model the neutron star via a polytropic equation of state, $P=K
\rho^{\Gamma}$ with $\Gamma=5/3$. For the following, we measure
distance and mass in units of the radius {\em R} and mass $M_{\rm NS}$
of the unperturbed (spherical) neutron star (13.4~km and
1.4~M$_{\odot}$ respectively), except where noted, so that the units
of time, density and velocity are
\begin{eqnarray}
\tilde{t}= 1.146 \times 10^{-4}~{\rm s} \times
\left(\frac{R}{13.4~\mbox{km}}\right)^{3/2}
\left(\frac{M_{\rm NS}}{1.4M_{\odot}}\right)^{-1/2}
\label {eq:deftunit}
\end{eqnarray}
\begin{eqnarray}
\tilde{\rho}= 1.14 \times 10^{18}~{\rm kg~m}^{-3} \times
\left(\frac{R}{13.4~\mbox{km}}\right)^{-3}
\left(\frac{M_{\rm NS}}{1.4M_{\odot}}\right)
\label{eq:defrhounit}
\end{eqnarray}
\begin{eqnarray}
\tilde{v}=0.39c \times \left(\frac{R}{13.4~\mbox{km}}\right)^{-1/2}
\left(\frac{M_{\rm NS}}{1.4M_{\odot}}\right)^{1/2}.
\label{eq:defvunit}
\end{eqnarray}
and we use corresponding units for derivative quantities such as
energy and angular momentum.
The black hole (of mass $M_{\rm BH}$) is modeled as a Newtonian point
mass, with a potential \( \Phi_{\rm BH}(r) = -GM_{\rm BH}/r \). We
model accretion onto the black hole by placing an absorbing boundary
at the Schwarzschild radius ($r_{Sch}=2GM_{\rm BH}/c^{2}$). Any
particle that crosses this boundary is absorbed by the black hole and
removed from the simulation. The mass and position of the black hole
are continously adjusted so as to conserve total mass and total
momentum.
Initial conditions corresponding to tidally locked binaries in
equilibrium are constructed in the co--rotating frame of the binary
for a range of separations {\em r} and a given value of the mass ratio
$q=M_{\rm NS}/M_{\rm BH}$~(Rasio \& Shapiro~1994; Paper~I). The binary
separation is defined henceforth as the distance between the black
hole and the center of mass of the SPH particles. During the
construction of these configurations, the specific entropies of all
particles are maintained constant, i.e. $K$=constant in
$P$=$K\rho^{\Gamma}$. The neutron star is modeled with $N=17,256$
particles in every case presented in this paper. To ensure uniform
spatial resolution, the masses of the particles were made proportional
to the Lane--Emden densities on the initial grid.
To carry out a dynamical run, the black hole and every particle are
given the azimuthal velocity corresponding to the equilibrium value of
the angular frequency $\Omega$ in an inertial frame, with the origin
of coordinates at the center of mass of the system. Each SPH particle
is assigned a specific internal enery
$u_{i}=K\rho^{(\Gamma-1)}/(\Gamma-1)$, and the equation of state is
changed to that of an ideal gas, where $P=(\Gamma-1)\rho u$. The
specific internal energy of each particle is then evolved according to
the first law of thermodynamics, taking into account the contributions
from the artificial viscosity present in SPH. During the dynamical
runs we calculate the gravitational radiation waveforms in the
quadrupole approximation.
We have included a term in the equations of motion that simulates the
effect of gravitational radiation reaction on the components of the
binary system. Using the quadrupole approximation, the rate of energy
change for a point--mass binary is given by (see Landau \&
Lifshitz~1975):
\begin{eqnarray}
\frac{dE}{dt}=-\frac{32}{5} \frac{G^{4}(M_{\rm NS}+M_{\rm BH})
(M_{\rm NS}M_{\rm BH})^{2}}{(cr)^{5}}
\label{eq:dedt}
\end{eqnarray}
and the rate of angular momentum loss by
\begin{eqnarray}
\frac{dJ}{dt}=-\frac{32}{5 c^{5}} \frac{G^{7/2}}{r^{7/2}}
M_{\rm BH}^{2} M_{\rm NS}^{2} \sqrt{M_{\rm BH}+M_{\rm NS}}.
\label{eq:djdt}
\end{eqnarray}
From these equations a radiation reaction acceleration for
each component of the binary can be obtained as
\begin{eqnarray}
\mbox{\boldmath $a$}^{*}=-\frac{1}{q(M_{\rm NS}+M_{\rm BH})}
\frac{dE}{dt}
\frac{\mbox{\boldmath $v$}^{*}}{(v^{*})^{2}}
\label{eq:reaction}\\
\mbox{\boldmath $a$}^{\rm BH}=-\frac{q}{M_{\rm NS}+M_{\rm BH}} \frac{dE}{dt}
\frac{\mbox{\boldmath $v$}^{\rm BH}}{(v^{\rm BH})^{2}}
\label{eq:reactionbh}
\end{eqnarray}
where $v^{*}$ is the velocity of the neutron star and $v^{\rm BH}$ that of
the black hole.
We have used this formula for our calculations to simulate the effect
of gravitational radiation reaction on the system. Clearly, the
application of equation~(\ref{eq:reactionbh}) to the black hole in our
calculations is trivial, since we always treat it as a point mass. For
the neutron star, we have chosen to apply the same acceleration to all
SPH particles. This value is that of the acceleration at the center of
mass of the SPH particles, so that equation~(\ref{eq:reaction}) now
reads:
\begin{eqnarray}
\mbox{\boldmath $a$}^{i}=-\frac{1}{q(M_{\rm NS}+M_{\rm BH})} \frac{dE}{dt}
\frac{\mbox{\boldmath $v$}^{*}_{cm}}{(v^{*}_{cm})^{2}},
\label{eq:reactionsph}
\end{eqnarray}
This formulation of the gravitational radiation reaction has been used
in SPH simulations by others~\cite{davies,zhuge,rosswog} in the case
of merging neutron stars, and it is usually switched off once the
stars come into contact, when the point--mass approximation clearly
breaks down. We are assuming then, that the polytrope representing the
neutron star can be considered as a point mass for the purposes of
including radiation reaction. If the neutron star is disrupted during
the encounter with the black hole, this radiation reaction must be
turned off, since our formula would no longer give meaningful
results. We have adopted a switch for this purpose, as follows: the
radiation reaction is turned off if the center of mass of the SPH
particles comes within a prescribed distance of the black hole
(effectively a tidal disruption radius). This distance is set to
$r_{tidal} = C R(M_{\rm BH}/M_{\rm NS})^{1/3}$, where {\em C} is a
constant of order unity.
\section{Results}\label{results}
We now describe our results. First, we present the initial conditions
that were used to perform the dynamical runs. We then describe the
general morphology of the coalescence events, the detailed structure
of the accretion disks that form as a result of the tidal disruption
of the neutron star, and the gravitational radiation signal.
\begin{table}
\caption{Basic parameters for each run~(Section 3.2)}
\label{parameters}
\begin{tabular}{@{}ccccccc}
Run & $q$ & $r_{RL}$ & $r_{i}$
& $t_{rad}$
& $t_{f}$ & $N$ \\
A & 1.00 & 2.67 & 2.70 & 30.51 & 200.0
& 17,256 \\
B & 0.80 & 2.81 & 2.85 & 29.97 & 200.0
& 17,256 \\
C & 0.31 & 3.59 & 3.60 & 25.69 & 200.0
& 17,256 \\
D & 0.31 & 3.59 & 3.60 & 15.97 & 200.0
& 17,256 \\
E & 0.10 & 5.01& 5.05 & 9.56 & 200.0
& 17,256 \\
\end{tabular}
\medskip
The table lists for each run the initial mass ratio, the orbital
separation at which the Roche lobe is filled, the initial orbital
separation, the time at which gravitational radiation reaction is
switched off in the simulation, and the initial number of particles.
\end{table}
\subsection{Evolution of the Binary}\label{initial}
To allow comparisons of results for differing equations of state, we
have run simulations with the same initial binary mass ratios as
previously explored~(Paper~I), namely $q$=1, $q$=0.8 and
$q$=0.31. Additionally we have examined the case with mass ratio
$q$=0.1. Equilibrium sequences of tidally locked binaries were
constructed for a range of initial separations, terminating at the
point where the neutron star overflows its Roche Lobe (at
$r=r_{RL}$). In Figure~\ref{jvsr}
\begin{figure*}
\psfig{width=\textwidth,file=jrA.eps,angle=0,clip=}
\caption{Total angular momentum {\em J} as a function of binary
separation {\em r} for various mass ratios. The solid line is the
result of the SPH calculation, the dashed line results from
approximating the neutron star as a compressible tri--axial ellipsoid
and the dotted line from approximating it as a rigid sphere.(a)~$q$=1;
(b)~$q$=0.8; (c)~$q$=0.31; (d)~$q$=0.1}
\label{jvsr}
\end{figure*}
we show the variation of total angular momentum {\em J} in these
sequences as a function of binary separation for the four values of
the mass ratio (solid lines). Following Lai, Rasio \&
Shapiro~\shortcite{LRSb}, we have also plotted the variation in {\em
J} that results from approximating the neutron star as compressible
tri--axial ellipsoid (dashed lines) and as a rigid sphere (dotted
lines).
In all cases, the SPH results for the $\Gamma=5/3$ polytrope are very
close to the ellipsoidal approximation until the point of Roche--Lobe
overflow. This result is easy to understand if one considers that the
softer the equation of state, the more centrally condensed the neutron
star is and the less susceptible to tidal deformations arising from
the presence of the black hole. For $\Gamma=3$~(Paper~I), the
variation in angular momentum as a function of binary separation was
qualitatively different (for high mass ratios) from our present
findings. For $q$=1 and $q$=0.8, total angular momentum attained a
minimum at some critical separation {\em before } Roche--Lobe overflow
occurred. This minimum indicated the presence of a dynamical
instability, which made the binary decay on an orbital timescale. This
purely Newtonian effect arose from the tidal interactions in the
system~\cite{LRSa}. In the present study, we expect all orbits with
initial separations $r\ge r_{RL}$ to be dynamically stable.
There is a crucial difference between the two polytropes considered in
Paper~I and here. For polytropes, the mass--radius relationship is \(
R \propto M^{(\Gamma-2)/(3\Gamma-4)}\). For $\Gamma$=3 this becomes
\(R \propto M^{1/5} \), while for $\Gamma$=5/3, \( R \propto
M^{-1/3}\). Thus, the polytrope considered in Paper~I responded to
mass loss by shrinking. The $\Gamma=5/3$ polytrope, considered here,
responds to mass loss by expanding, as do neutron stars modeled with
realistic equations of state~\cite{arnett}--the dynamical disruption
of the star reported below seems to be related to this effect. For the
polytropic index considered in Paper~I, the star was not disrupted
(see also Lee \& Klu\'{z}niak 1995; 1997; Klu\'{z}niak \& Lee 1998),
but we find no evidence in any of our dynamical calculations for a
steady mass transfer in the binary, such as the one suggested in the
literature (e.g. Blinnikov et al.~1984; Portegies Zwart~1998).
Using equations~(\ref{eq:dedt}) and~(\ref{eq:djdt}) one can compute
the binary separation as a function of time for a point--mass binary
in the quadrupole approximation, and obtain
\begin{eqnarray}
r=r_{i}\left( 1-t/t_{0} \right)^{1/4}, \label{eq:ptdecay}
\end{eqnarray}
with \( t_{0}^{-1}=256 G^{3}M_{\rm BH} M_{\rm NS} (M_{\rm BH}+M_{\rm
NS})/(5r_{i}^{4}c^{5})\). Here $r_{i}$ is the separation at $t$=0. For
black hole--neutron star binaries studied in this paper, the timescale
for orbital decay because of angular momentum loss to gravitational
radiation, $t_{0}$, is on the order of the orbital period, $P$ (for
$q$=1, at an initial separation $r_{i}$=2.7 we find $t_{0}=56.81
\times \tilde{t}$=6.5~ms) and $P=19.58 \times \tilde{t}$=2.24~ms), so
one must analyze whether hydrodynamical effects will drive the
coalescence on a comparable timescale. We have performed a dynamical
simulation for $q$=1, with an initial separation on the verge of Roche
Lobe overflow, at $r$=2.7, {\em without} including radiation reaction
in the equations of motion. We show in Figure~\ref{rvst}a
\begin{figure*}
\psfig{width=\textwidth,file=rtA.eps,angle=0,clip=}
\caption{(a) Binary separation as a function of time for the test run
with mass ratio $q=1$ {\em without} gravitational radiation reaction
(solid line), and for a point--mass binary with the same initial
separation emitting gravitational waves, calculated in the quadrupole
approximation (dashed line).(b) Binary separation for runs A, C, D and
E (Table~\ref{parameters}) and for the corresponding point--mass
binaries. The vertical line across the solid curves marks the time
$t=t_{rad}$. The curves terminate when the axis ratio of the deformed
neutron star (in the orbital plane) is $a_{2}/a_{1} \approx 2$.}
\label{rvst}
\end{figure*}
the binary separation as a function of time for this calculation
(solid line) as well as the separation for a point--mass binary
decaying in the quadrupole approximation, using
equation~(\ref{eq:ptdecay}) (dashed line). In the dynamical simulation
the orbital separation remains approximately constant, and begins to
decay rapidly around $t$=110 (in the units defined in
equation~[\ref{eq:deftunit}]), when mass loss from the neutron star
becomes important. Clearly at this stage hydrodynamical effects are
dominant, but one must include radiation reaction in the early stages
of the process. There is an added (practical) benefit derived from
including radiation reaction in these calculations. As seen in
Figure~\ref{rvst}a, it takes a full 15~ms for the orbit to become
unstable. Simulating the behavior of the system at high resolution
(practically no SPH particles have been accreted at this stage) for
such a long time is computationally expensive, whereas accretion in
the early stages of the simulation allows us to perform, in general,
more calculations at higher resolution. We have thus included
radiation reaction in all the runs (A through E) presented in this
paper, and adopted a switch as described in section~\ref{method}.
\begin{figure*}
\psfig{width=\textwidth,file=rhocontourstrad.eps,angle=0,clip=}
\caption{Density contours in the orbital plane at $t \approx t_{rad}$,
i.e., when the gravitational radiation reaction is switched off for
(a) run A, (b) run C, (c) run D, (d) run E. All contours are
logarithmic and equally spaced every 0.5 dex. The lowest (outermost)
contour is at $\log{\rho}=-3$ (in the units defined in
equation~[\ref{eq:defrhounit}]), and bold contours are plotted at
$\log{\rho}=-3,-2,-1,0$ (if present). The center of mass of the system
is at the origin of the inertial coordinate frame, all distances are
in units of the initial radius of the unperturbed star. The dark disk
of radius $r_{Sch}$ represents the black hole.}
\label{rhocontourstrad}
\end{figure*}
\begin{figure*}
\psfig{width=\textwidth,file=rhocontoursq1.eps,angle=0,clip=}
\caption{Density contours after the disruption of the polytrope for
$q=1$ (Run A) in the equatorial plane (left panels), and in the
meridional plane containing the black hole (right panels). All
contours are logarithmic and equally spaced every 0.5 dex. (a)--(b)
The lowest contour is at $\log{\rho}=-5$ (in the units defined in
equation~[\ref{eq:defrhounit}]), and bold contours are plotted at
$\log{\rho}=-5,-4,-3,-2$. (c)--(d) The lowest contour is at
$\log{\rho}=-6$ (in the units defined in
equation~[\ref{eq:defrhounit}]), and bold contours are plotted at
$\log{\rho}=-6,-5,-4$.}
\label{rhocontoursq1}
\end{figure*}
\begin{figure*}
\psfig{width=\textwidth,file=rhocontoursq031b.eps,angle=0,clip=}
\caption{Same as Figure~\ref{rhocontoursq1} but for $q=0.31$ (Run
D). All contours are logarithmic and equally spaced every 0.5
dex. (a)--(d) The lowest contour is at $\log{\rho}=-6$ (in the units
defined in equation~[\ref{eq:defrhounit}]), and bold contours are
plotted at $\log{\rho}=-6,-5,-4,-3$ (if present).}
\label{rhocontoursq031b}
\end{figure*}
\begin{figure*}
\psfig{width=\textwidth,file=rhocontoursq01.eps,angle=0,clip=}
\caption{Same as Figure~\ref{rhocontoursq1} but for $q=0.1$ (Run
E). All contours are logarithmic and equally spaced every 0.5
dex. (a)--(b) The lowest contour is at $\log{\rho}=-6$ (in the units
defined in equation~[\ref{eq:defrhounit}]), and bold contours are
plotted at $\log{\rho}=-6,-5,-4,-3$. (c)--(d) The lowest contour is at
$\log{\rho}=-6.5$ (in the units defined in
equation~[\ref{eq:defrhounit}]), and bold contours are plotted at
$\log{\rho}=-6,-5$.}
\label{rhocontoursq01}
\end{figure*}
\begin{figure*}
\psfig{width=\textwidth,file=mdotmbhA.eps,angle=0,clip=}
\caption{Mass accretion rate onto the black hole~(a) and mass of the
black hole~(b) for all runs. In both panels, the curves for runs C and
D coincide until $t$=15.97. $M_{\rm BH}$ is in units of
1.4~M$_{\odot}$ and time is in units of $\tilde{t}$ (see
equation~[\ref{eq:deftunit}]).}
\label{mdot}
\end{figure*}
\subsection{Run parameters}\label{param}
In Table~\ref{parameters} we present the parameters distinguishing
each dynamical run we performed. All times are in units of
$\tilde{t}$~(equation~[\ref{eq:deftunit}]) and all distances in units
of {\em R}, the unperturbed (spherical) stellar radius. The runs are
labeled with decreasing mass ratio (increasing black hole mass), from
$q$=1 down to $q$=0.1. All simulations were run for the same length of
time, $t_{final}=200$, equivalent to 22.9~ms (this covers on the order
of ten initial orbital periods for the mass ratios considered).
The fifth column in Table~\ref{parameters} shows the value of
$t_{rad}$, when radiation reaction is switched off according to the
criterion described in section~\ref{method}. In
Figure~\ref{rhocontourstrad} we show density contours in the orbital
plane for runs A, C, D and E at times very close to $t=t_{rad}$. The
corresponding plot for run B is very similar to that for run A. Note
that runs C and D differ {\em only} in the corresponding value of
$t_{rad}$. For run D there is little doubt that approximating the
neutron star as a point--mass is still reasonable at this stage, while
for run C this is clearly not the case. We can then use these two runs
to gauge the effect of our simple radiation reaction formulation on
the outcome of the coalescence event. We note here that run E is
probably beyond the limit of what should be inferred from a Newtonian
treatment of such a binary system. The black hole is very large
compared to the neutron star, and the initial separation
($r_{i}=5.05$, equivalent to 67.87~km) is such that the neutron star
is within the innermost stable circular orbit of a test particle
around a Schwarzschild black hole of the same mass ($r_{ms}=9.17$,
equivalent to 123.26~km). Thus we present in Appendix~\ref{pw} a
dynamical run with initial mass ratio $q=0.1$ making use of a
pseudo--Newtonian potential for the black hole. For the following, we
will at times omit a discussion of run B, as it is qualitatively and
quantitatively very similar to run A (both of these have a relatively
high mass ratio).
\subsection{Morphology of the mergers}
The initial configurations are close to Roche Lobe overflow, and mass
transfer from the neutron star onto the black hole starts within one
orbital period for all runs, A through E. Once accretion begins, the
total number of particles decreases. Since this compromises
resolution, we made a modification to the code for run E to avoid the
number of particles from dropping below 9,000. This is done simply by
splitting a given fraction of the particles $N_{split}$ and creating
$2N_{split}$ particles from them. Total mass and momentum are
conserved during this procedure, and it can be shown that the
numerical noise introduced into the smoothed density $\langle\rho
\rangle$ by doing this is of the order of the accuracy of the SPH
method itself, $\mathcal{O}$($h^{2}$), where $h$ is the smoothing
length~\cite{meglicki}.
In every run the binary separation (solid lines in in
Figure~\ref{rvst}b) initially decreases due to gravitational radiation
reaction. For high mass ratios, (runs A, B) the separation decays
faster that what would be expected of a point--mass binary. This is
also the case for a stiff equation of state, in black hole--neutron
star mergers~(Paper~I) as well as in binary neutron star
mergers~\cite{RS94}, and merely reflects the fact that hydrodynamical
effects are playing an important role. For the soft equation of state
studied here, there is the added effect of `runaway' mass transfer
because of the mass--radius relationship (see
section~\ref{initial}). For runs C and D, the solid and dashed lines
in Figure~\ref{rvst}b follow each other very closely, indicating that
the orbital decay is primarily driven by angular momentum losses to
gravitational radiation. For run E, the orbit decays more slowly than
what one would expect for a point--mass binary. This is explained by
the fact that there is a large amount of mass transfer (10\% of the
initial neutron star mass has been accreted by $t=t_{rad}$ in this
case) in the very early stages of the simulation, substantially
altering the mass ratio in the system (the dashed curves in
Figure~\ref{rvst} are computed for a fixed mass ratio). From the
expression for the timescale for orbital decay $t_{0}$ in
equation~(\ref{eq:ptdecay}), it is apparent that at constant total
mass, lowering the mass ratio slows the orbital decay, when $q<0.5$.
The general behavior of the system is qualitatively similar for every
run. Figures~\ref{rhocontoursq1},~\ref{rhocontoursq031b} and
~\ref{rhocontoursq01} show density contours in the orbital plane (left
columns) and in the meridional plane containing the black hole (right
columns) for runs A, D and E respectively, at $t=50$ and $t=t_{f}=200$
(equivalent to 5.73~ms and 22.9~ms). The corresponding plots for runs
B and C are very similar to those for runs A and D, respectively. The
neutron star becomes initially elongated along the binary axis and an
accretion stream forms, transferring mass to the black hole through
the inner Lagrange point. The neutron star responds to mass loss and
tidal forces by expanding, and is tidally disrupted. An accretion
torus forms around the black hole as the initial accretion stream
winds around it. A long tidal tail is formed as the material furthest
from the black hole is stripped from the star. Most of the mass
transfer occurs in the first two orbital periods and peak accretion
rates reach values between 0.04 and 0.1---equivalent to 0.49 and 1.22
M$_{\odot}$/ms (see Figure~\ref{mdot}). The mass accretion rate then
drops and the disk becomes more and more azimuthally symmetric,
reaching a quasi--steady state by the end of the simulations.
We show in Figure~\ref{energies} the various energies of the system
(kinetic, internal, gravitational potential and total) for runs A, D
and E.
\begin{figure*}
\psfig{width=\textwidth,file=energiesA.eps,angle=0,clip=}
\caption{(a) Various energies of the system as a function of time for
run A, (b) energies in run D, (c) energies in run E. The kinetic (T),
internal (U), gravitational potential (W) and total (E) energies are
indicated in units of $3.8\times 10^{53}$~erg. (d) Angular momentum as
a function of time for every run in units of $4.37\times
10^{42}$~kg~m$^{2}$~s$^{-1}$.}
\label{energies}
\end{figure*}
The dramatic drop in total internal energy reflects the intense mass
accretion that takes place within the first couple of
orbits. Figure~\ref{energies} also shows [in panel (d)] the total
angular momentum of the system (the only contribution to the total
angular momentum not plotted is the spin angular momentum of the black
hole, see below). Angular momentum decreases for two reasons. First,
if gravitational radiation reaction is still acting on the system, it
will decrease approximately according to
equation~(\ref{eq:djdt}). Second, whenever matter is accreted by the
black hole, the corresponding angular momentum is removed from our
system. In reality, the angular momentum of the accreted fluid would
increase the spin of the black hole. We keep track of this accreted
angular momentum and exhibit its value in Table~2 as the Kerr
parameter of the black hole. This shows up as a decrease in the total
value of {\em J}. It is clear from runs C and D that the value of
$t_{rad}$ influences the peak accretion rate and the mass of the black
hole (particularly immediately after the first episode of heavy mass
transfer). The maximum accretion rate is different for run C and D by
about a factor of 1.4. This is easy to understand, since radiation
reaction increases angular momentum losses, and hence more matter is
accreted per unit time when it is present.
\subsection{Accretion disk structure}
In Table~\ref{disks} we show several parameters pertaining to the
final accretion structure around the black hole for every run. The
disk settles down to a fairly azimuthally symmetric structure within a
few initial orbital periods (except for the long tidal tail, which
always persists as a well--defined structure), and there is a
baryon--free axis above (and below) the black hole in every case
(Figure~\ref{mtheta}). We have calculated the mass of the remnant
disk, $M_{disk}$, by searching for the amount of matter that has
sufficient specific angular momentum {\em j} at the end of the
simulation to remain in orbit around the black hole (as in Ruffert \&
Janka~1999). This material has $j>j_{crit}=\sqrt{6}GM_{t}/c$, where
$M_{t}$ is the total mass of the system. The values shown in
Table~\ref{disks} are equivalent to a few tenths of a solar mass, and
again the effect of $t_{rad}$ can be seen by comparing runs C and D,
where the disk masses differ by a factor of 1.2. By the end of the
simulations, between 70\% and 80\% of the neutron star has been
accreted by the black hole. It is interesting to note that the {\em
final} accretion rate (at $t=t_{f}$) appears to be rather insensitive
to the initial mass ratio, and is between $2\times 10^{-4}$ and
$5\times 10^{-4}$ (equivalent to 2.4 and 6.1 M$_{\odot}$~s$^{-1}$
respectively). From this final accretion rate we have estimated a
typical timescale for the evolution of the accretion disk,
$\tau_{disk}=M_{disk}/\dot{M}_{final}$---for reference, $\tau=100$, in
the units of equation~(\ref{eq:deftunit}) corresponds to $11.5$~ms and
thus the values of $\tau_{disk}$ given in Table~\ref{disks} are
between 47 and 63~ms. Despite the difference in the initial mass
ratios and the typical sizes of the disks ($r_{0}$ is the radial
distance from the black hole to the density maximum at $t=t_{f}$), the
similar disk masses and final accretion rates make the lifetimes
comparable for every run.
We have plotted azimuthally averaged density and internal energy
profiles in Figure~\ref{diskprofiles} for runs A, D and E. The
specific internal energy is greater towards the center of the disk,
and flattens out at a distance from the black hole roughly
corresponding the density maximum, at $u\simeq 2 \times 10^{-2}$. This
value corresponds to $2.74 \times 10^{18}$~erg~g$^{-1}$ or
2.9~MeV/nucleon and is largely independent of the initial mass
ratio. The inner regions of the disks have specific internal energies
that are greater by approximately one order of magnitude.
Additionally, panel (d) in the same figure shows the azimuthally
averaged distribution of specific angular momentum {\em j} in the
orbital plane for all runs. The curves terminate at $r_{in}=2r_{Sch}$.
Pressure support in the inner regions of the accretion disks makes the
rotation curves sub--Keplerian, while the flattening of distribution
marks the outer edge of the disk and the presence of the long tidal
tail (see Figure~\ref{ejected}),which has practically constant
specific angular momentum.
The Kerr parameter of the black hole, given by $a=J_{\rm BH}c/G M_{\rm
BH}^{2}$, is also shown in Table~\ref{disks}. We have calculated it
from the amount of angular momentum lost via accretion onto the black
hole (see Figure~\ref{energies}d), assuming that the black hole is
{\em not} rotating at $t=0$. The final specific angular momentum of
the black hole is smaller for lower mass ratios simply because the
black hole is initially more massive when {\em q} is smaller. The
difference in the value of {\em a} for runs C and D is important
(almost a factor of 2), and again reflects the influence of
gravitational radiation reaction (for a larger value of $t_{rad}$ the
black hole is spun up to a greater degree because of the larger amount
of accreted mass).
It is of crucial importance for the production of GRBs from such a
coalescence event that there be a baryon--free axis in the system
along which a fireball may expand with ultrarelativistic
velocities~(M\'{e}sz\'{a}ros \& Rees 1992; 1993). We have calculated
the baryon contamination for every run as a function of the
half--angle $\Delta \theta$ of a cone directly above (and below) the
black hole and along the rotation axis of the binary that contains a
given amount of mass $\Delta M$. Table~\ref{disks} shows these angles
(in degrees) for $\Delta M=10^{-3},10^{-4},10^{-5}$ (equivalent to
$1.4\times 10^{-3},1.4\times 10^{-4},1.4\times 10^{-5}$~M$_{\odot}$
respectively). There is a greater amount of pollution for high mass
ratios (the disk is geometrically thicker compared to the size of the
black hole), but in all cases only modest angles of collimation are
required to avoid contamination. We note here that the values for
$\theta_{-5}$ are rough estimates at this stage since they are at the
limit of our numerical resolution in the region directly above the
black hole. This can be seen by inspection in Figure~\ref{mtheta}
\begin{figure}
\psfig{width=85mm,file=mthetaD.eps,angle=0,clip=}
\caption{Enclosed mass as a function of half--angle $\Delta \theta$
(measured from the rotation axis in degrees) for all runs at
$t=t_{f}$. The mass resolution varies from approximately $5 \times
10^{-6}$ to $5 \times 10^{-5}$ (corresponding to $7 \times
10^{-6}$~M$_{\odot}$ and $7 \times 10^{-5}$~M$_{\odot}$ respectively)
in the region directly above the black hole.}
\label{mtheta}
\end{figure}
where we show the enclosed mass as a function of half--angle $\Delta
\theta$ for all runs at $t=t_{f}$.
\begin{figure*}
\psfig{width=\textwidth,file=diskkepA.eps,angle=0,clip=}
\caption{Azimuthally averaged profiles for the density, $\rho$, and
the specific internal energy, {\em u} ($u/1000$ is plotted), of the
accretion torus at $t=t_{f}$ for runs A~(a), D~(b) and E~(c). The
inner edge of the curves is at $r=2 r_{Sch}$. At this stage in the
simulation the tori are close to being azimuthally symmetric. For
reference, $\rho=10^{-4}$ corresponds to $1.14 \times
10^{11}~\mbox{g~cm$^{-3}$}$ and $u/1000=10^{-4}$ corresponds to $1.37
\times 10^{19}~\mbox{erg~g$^{-1}$}$. Panel~(d) shows the distribution
of specific angular momentum {\em j} for all runs at $t=t_{f}$ (solid
lines, A, B, C, D, E) and the specific angular momentum of a Keplerian
accretion disk for the same black hole mass (dashed lines, {\it A},
{\it B}, {\it C}, {\it D}, {\it E}).}
\label{diskprofiles}
\end{figure*}
\begin{table*}
\caption{Accretion disk structure}
\label{disks}
\begin{tabular}{@{}cccccrrcccccc}
Run & $q$ & $M_{disk}$ & $M_{acc}$ & $\dot{M}_{max}$
& $\dot{M}_{final}$
& $M_{ejected}$
& $r_{0}$ & $\tau_{disk}$ & $J_{\rm BH}c/G M_{\rm BH}^{2}$ &
$\theta_{-3}$
& $\theta_{-4}$ & $\theta_{-5}$ \\
A & 1.00 & 0.188 & 0.78 & 0.068 & $4 \cdot 10^{-4}$ & $9.51 \cdot 10^{-3}$
& 4.83 & 472
& 0.517 & 20 & 12 & 8 \\
B & 0.80 & 0.198 & 0.77 & 0.060 & $5 \cdot 10^{-4}$ & $5.41 \cdot 10^{-3}$
& 4.46 & 409
& 0.497 & 25 & 10 & 3 \\
C & 0.31 & 0.184 & 0.79 & 0.062 & $3.5 \cdot 10^{-4}$ & $3.95 \cdot 10^{-3}$
& 6.32 & 532
& 0.313 & 37 & 27 & 22 \\
D & 0.31 & 0.226 & 0.74 & 0.045 & $4 \cdot 10^{-4}$ & $16.12 \cdot 10^{-3}$
& 7.44 & 551
& 0.173 & 40 & 21 & 10 \\
E & 0.10 & 0.072 & 0.86 & 0.095 & $2 \cdot 10^{-4}$ & $3.18 \cdot 10^{-3}$
& 10.41 & 410
& 0.114 & 52 & 42 & 32 \\
\end{tabular}
\medskip
In the last three columns, $\theta_{-n}$ is the half--angle of a cone
above the black hole and along the rotation axis of the binary that
contains a mass $M=10^{-n}$.
\end{table*}
\subsection{Ejected mass and r--process}
To calculate the amount of dynamically ejected mass during the
coalescence process, we look for matter that has a positive total
energy (kinetic+gravitational potential+internal) at the end of each
simulation. Figure~\ref{ejected} shows a large--scale view of the
system at $t=t_{f}$ for runs D and E. The thick black line running
across the tidal tail in each case divides matter that is bound to the
black hole from that which may be on outbound trajectories. This
matter comes from the part of the neutron star that was initially
furthest from the black hole and was ejected through the outer
Lagrange point in the very early stages of mass transfer. We find that
a mass between $4.4 \times 10^{-3}$~M$_{\odot}$ and $2.2 \times
10^{-2}$~M$_{\odot}$ can potentially be ejected in this fashion (see
Table~\ref{disks}). This is very similar to what has recently been
calculated for binary neutron star mergers for a variety of initial
configurations~\cite{rosswog}. Since it is believed that the event
rate for binary neutron star mergers is comparable to that of black
hole--neutron star mergers, this could prove to be a sizable
contribution to the amount of observed r--process material. This
result appears to be strongly dependent on the equation of state,
since we previously observed no significant amount of matter being
ejected for a system with a stiff equation of state~(Paper~I). For
binary neutron star mergers, Rosswog et al.~1999 have obtained the
same qualitative result. We also note that there is a significant
difference in the amount of ejected mass for runs C and D
(approximately a factor of four), due to the difference in the values
of $t_{rad}$ (in run D the system loses less angular momentum and thus
matter escapes with greater ease).
\begin{figure*}
\psfig{width=\textwidth,file=rhocontoursxy70largeA.eps,angle=0,clip=}
\caption{Final density contours in the orbital plane for runs D (left
panel) and E (right panel). All contours are logarithmic and equally
spaced every 0.5 dex. (a) The lowest contour is at $\log{\rho}=-7.5$
(in the units defined in equation~[\ref{eq:defrhounit}]), and bold
contours are plotted at $\log{\rho}=-7,-6,-5,-4$. (b) The lowest
contour is at $\log{\rho}=-8$ (in the units defined in
equation~[\ref{eq:defrhounit}]), and bold contours are plotted at
$\log{\rho}=-8,-7,-6,-5$. The thick black line running across the
tidal tail in each frame divides the matter that is bound to the black
hole from that which is on outbound trajectories.}
\label{ejected}
\end{figure*}
\subsection{Gravitational radiation waveforms and luminosities}
\begin{figure*}
\psfig{width=\textwidth,file=gradA.eps,angle=0,clip=}
\caption{(a)--(c):Gravitational radiation waveforms (one polarization)
for runs A~(a), D~(b) and E~(c). The solid lines are the results from
the dynamical simulations, while the dashed lines show the emission
from a point--mass binary with the same initial separation and mass
ratio, calculated in the quadrupole approximation.(d) Gravitational
radiation luminosity. All quantities are plotted in geometrized units
such that $G$=$c$=1 ($L_{0}=c^{5}/G=3.64 \times
10^{59}\mbox{erg~s$^{-1}$}$).}
\label{gw}
\end{figure*}
\begin{table*}
\caption{Gravitational radiation}
\label{waves}
\begin{tabular}{@{}ccccc}
Run & $q$ & $(r_{0}R/M_{\rm NS}^{2})h_{max}$
& $(R/M_{\rm NS})^{5}(L_{max}/L_{0})$
& $(R^{7/2}/M_{\rm NS}^{9/2})\Delta E_{GW}$ \\
& & & & \\
A & 1.00 & 1.93 & 0.37 & 6.53 \\
B & 0.80 & 2.26 & 0.44 & 8.24 \\
C & 0.31 & 4.40 & 1.45 & 21.11 \\
D & 0.31 & 4.46 & 1.50 & 20.77 \\
E & 0.10 & 9.42 & 5.90 & 60.08 \\
\end{tabular}
\medskip
All quantitites are given in geometrized units such that $G=c=1$, and
$L_{0}=c^{5}/G=3.64\times 10^{59}$~erg~s$^{-1}$.
\end{table*}
The emission of gravitational radiation is calculated in all our
models in the quadrupole approximation (see e.g. Finn~1989; Rasio \&
Shapiro~1992), and can be obtained directly from the hydrodynamical
variables of the system. The calculation of the gravitational
radiation luminosity then requires only an additional numerical
differentiation. Figure~\ref{gw} shows the computed waveforms and
luminosities, along with what the waveforms would be for a point--mass
binary, also calculated in the quadrupole approximation. It is
apparent that hydrodynamical effects play an important role
particularly for high mass ratios in the early stages of the
coalescence process (see panel~(a) in Figure~\ref{gw}). When the
neutron star is tidally disrupted, the amplitude of the waveform drops
abruptly and practically to zero, since a structure that is almost
azimuthally symmetric has formed around the black hole. This is in
stark contrast to what occurred for a stiff equation of
state~(Lee~1998; Paper~I; Klu\'{z}niak \& Lee~1998), when the binary
system survived the initial episode of mass transfer and a stable
binary was the final outcome. In fact, these waveforms resemble more
the case of a double neutron star merger with a soft equation of
state~\cite{RS92}, in which the coalescence resulted in a compact,
azimuthally symmetric object surrounded by a dense halo and spiral
arms. Table~\ref{waves} shows the maximum amplitude $h_{max}$ for an
observer located a distance $r_{0}$ away from the system along the
axis of rotation, the maximum luminosity $L_{max}$ and the total enery
$\Delta E_{GW}$ emitted during the event. This last number should be
taken only as an order of magnitude estimate since it depends on the
choice of the origin of time. The peak luminosities are $(R/M_{\rm
NS})^{5}(L_{max}/L_{0})=0.37$ for run A, $(R/M_{\rm
NS})^{5}(L_{max}/L_{0})=1.50$ for run D and $(R/M_{\rm
NS})^{5}(L_{max}/L_{0})=5.90$ for run E (equivalent to $ 1.12 \times
10^{55}$erg~s$^{-1}$, $4.55 \times 10^{55}$erg~s$^{-1}$ and $ 1.79
\times 10^{56}$erg~s$^{-1}$ respectively). We note that although the
waveforms for runs C (not plotted) and D (panel (b) in
Figure~\ref{gw}) are very similar, the maximum amplitudes and the
luminosity differ by about 1.3\% and 3.4\% respectively, a small but
non--negligible amount. This is again a reflection of the way in which
the radiation reaction was formulated, and indicates that a more
rigorous treatment of this effect is necessary.
\section{Summary and Discussion}\label{discussion}
We have presented results of hydrodynamical simulations of the binary
coalescence of a black hole with a neutron star. We have used a
polytropic equation of state (with index $\Gamma=5/3$) to model the
neutron star, and a Newtonian point mass with an absorbing surface at
the Schwarzschild radius to represent the black hole. All our
computations are strictly Newtonian, but we have included a term that
approximates the effect of gravitational radiation reaction in the
system. We have also calculated the emission of gravitational
radiation in the quadrupole approximation.
We have found that for every mass ratio investigated ($M_{\rm
NS}/M_{\rm BH}$=1, 0.8, 0.31 and 0.1) the $\Gamma=5/3$ polytrope
(`neutron star') is entirely disrupted by tidal forces, and a dense
accretion torus, containing a few tenths of a solar mass, forms around
the black hole. The maximum densities and specific internal energies
in the tori are on the order of $10^{11}~\mbox{g~cm$^{-3}$}$ and
$10^{19}~\mbox{erg~g$^{-1}$}$ (or 10~MeV/nucleon) respectively (all
simulations were run for approximately 22.9~ms). The final accretion
rate is between 2 and 6 solar masses per second, and hence the
expected lifetime of the torus $\tau_{disk}=M_{disk}/\dot{M}_{final}$
is between 40 and 60 milliseconds.
The rotation axis of the system remains free of matter to a degree
that would not hinder the production of a relativistic fireball,
possibly giving rise to a gamma ray burst. Although the duration of
such a bursts would still be too short to power the longest GRBs, the
present scenario could well account for the subclass of short
bursts~\cite{kouveliotou}. A significant amount of matter (between
$10^{-2}$ and $10^{-3}$ solar masses) is dynamically ejected from the
system, and could contribute significantly to the observed abundances
of r--process material in our galaxy. The gravitational radiation
signal is very similar to that of a point--mass binary until the
beginning of mass transfer, particularly for low mass ratios. After
mass transfer starts, the amplitude of the waveforms drops
dramatically on a dynamical timescale when the accretion torus is
formed. In every aspect, the results are dramatically different from
what occurs for a stiff equation of state.
\section*{Acknowledgments}
We gratefully acknowledge financial support from DGAPA--UNAM and KBN
grant 2P03D01311. W.L. thanks Craig Markwardt for helpful discussions
concerning the effect of a soft equation of state on the system. It is
a pleasure to thank the referee for his most helpful comments.
|
\section{Introduction}
CP violation was first observed in kaon decays over 30 years ago.
In the Standard Model the Cabibbo-Kobayashi-Maskawa (CKM) weak
and mass eigenstates mixing matrix can provide a possible mechanism
for explanation of the observed CP violation effects.
The unitary CKM matrix is described by four physical parameters,
one of them being a complex phase.
An analysis of unitarity constraints in which
all of the elements are of the same order of magnitude,
e.g.:
$V_{ud}V_{ub}^* + V_{cd}V_{cb}^* + V_{td}V_{tb}^* = 0$
and
$V_{ud}V_{td}^* + V_{us}V_{ts}^* + V_{ub}V_{tb}^* = 0$
\noindent
provides a rudimentary test of the CKM description of CP violation.
The magnitude of the complex elements have been determined
from B-hadron lifetimes, branching fractions and
- more recently - precise flavor oscillations measurements.
The relative complex phases of the CKM matrix elements can be studied
in measurements of the CP asymmetries in B-decays.
An analysis of the asymmetry in the decay rates of $B^0$ and $\bar{B}^0$
to a common CP eigenstate $\JKS$
provides a measurement of the
phase $\beta \equiv arg(- \frac{ V_{cd}V_{cb}^*}{V_{td}V_{tb}^*} )$.
The asymmetry, ${\cal{A}}_{CP} \equiv \frac{N(\bar{B}^0) - N(B^0)}
{N(\bar{B}^0) + N(B^0)}$, where $N(\bar{B}^0)$ and $N(B^0)$
are numbers of observed decays to $\JKS$ given the known flavor of
the B meson at production, arises from the interference between
the direct decay path,
$\bar{B^0} \rightarrow J/\psi K^0_s$, and the mixed decay path,
$\bar{B^0} \rightarrow B^0 \rightarrow J/\psi K^0_s$.
The CP asymmetry ${\cal{A}}_{CP}$ depends on the CP phase difference
between the two amplitudes, $\beta$
and the flavor oscillations term represented
by $\sin(\Delta m_d t)$, where $\Delta m_d$ is the mass difference
between the two $B_d^0$ mass eigenstates, and {\it t} is proper decay time.
In the Standard Model ${\cal{A}}_{CP} \simeq \sin(2\beta) \sin(\Delta m_d t)$
since other contributions are expected to be very small.
Values of $\sin(2\beta)$ are
constrained to a range of $0.3 \leq \sin(2\beta) \leq 0.9$ from
indirect electroweak measurements.\cite{Mele} Last year
the OPAL collaboration at LEP reported \cite{Opal}
$\sin(2\beta)=4\pm2\pm1$, using
a sample of 14 $B^0/\bar{B^0} \rightarrow J/\psi K^0_s$ decays.
In this talk I will describe the flavor tagging techniques adapted
by CDF for application in the hadron collider
environment and discuss their performance
in the flavor oscillation measurements.
I will also report on the CP analysis of
$\B0JKS$ decays reconstructed in a data sample of
110 pb$^{-1}$ collected by the CDF detector at the Tevatron collider
at Fermilab. The description of the CDF detector
can be found in previous publications \cite{cdfd1,cdfd2}.
\section{Data Sample}
The reconstruction of $B^0_d$ mesons was done via the
decay $\BJKS$, with $J/\psi \rightarrow \mu^+ \mu^-$ and
$K^0_s \rightarrow \pi^+ \pi^-$.
The selection of the $B$ candidates begins by identifying $\jpsi$
particles that decay into two muons of opposite charge.
\begin{figure}[ht]
\centerline{
\epsfysize 6.0cm
\epsffile{k0ssvxall_desc.eps}
\epsfysize 6.0cm
\epsffile{k0sctcall_desc.eps}}
\caption{Normalized mass distributions of the $\B0JKS$ candidates.
\label{fig:jpsiks}}
\end{figure}
\noindent
All pairs of the oppositely charged particle tracks are considered to be
candidates for the $K^0_s$ decay products.
The B candidate mass and momentum are calculated subject to the
constraints that the invariant masses of the muon
pair and the pion pair are equal to the world average mass of their parent
particle, $\jpsi$ and $K^0_s$, respectively;
come from separate vertices; the reconstructed $K^0_s$
candidate points back to the $\jpsi$ vertex; and the $\JKS$ system
points back to the primary interaction vertex.
The silicon micro-vertex
detector (SVX) information was used for these constraints when available.
For a B candidate with both muons measured in the silicon vertex detector,
the typical mass resolution is $\sim 10~MeV/c^2$,
and the proper decay length resolution is $\sim 50~\mu m$.
The normalized mass distribution,
$M_N = (m_{\mu \mu \pi \pi} - M_{PDG})/\sigma_{fit}$,
is shown in figure ~\ref{fig:jpsiks}. The total number
of reconstructed B mesons is $395\pm31$, with a signal-to-noise
ratio of 0.7. The sample with both muons reconstructed in SVX contains
202$\pm$18 events with a signal-to-noise ratio of 0.9, and the remainder
of the sample contains 193$\pm$26 events with a signal-to-noise ratio of 0.5.
\vspace*{-0.2cm}
\section{Identification of B Flavor at Production and Decay}
\subsection{Opposite Side Tagging with Soft Lepton and Jet Charge}\label{subsec:slt}
The Opposite Side Flavor Tagging techniques
use a triggered lepton and a reconstructed secondary vertex to identify
the flavor of the B meson at the decay time. The flavor of the B
meson at the production time is determined either from the charge
of the jet on the side opposite the triggered lepton or by
the presence of another lepton in the event.
These flavor tagging methods were studied using high statistics samples
of semileptonic B decays,\cite{cdfsltjqc,lepd} as illustrated in
figure ~\ref{fig:bbar}.
The Opposite Side Flavor tagging algorithms were calibrated using a sample
of $\sim 1,000$ $B_u^{\pm} \rightarrow J/\psi K^{\pm}$ decays.
\vspace{-0.3cm}
\begin{figure}[ht]
\centerline{
\epsfysize 6.0cm
\epsffile{sst_plot.ps}
\epsfysize 6.0cm
\epsffile{ssall.eps}}
\caption{Asymmetry as a function of the proper decay length ct.
Left: Same Side tagging applied to $B \rightarrow \ell D^{*}$;
Right: Soft lepton and Jet Charge flavor tagging.
Results from an unbinned likelihood fit are
superimposed on the data points. \label{fig:bbar}}
\end{figure}
The Soft Lepton Tagging (SLT) algorithm correlates the charge of
the second lepton in the event with the flavor of the B at the
production time.
Its performance was checked through observation of the
$B_d^0-\overline{B_d^0}$ flavor oscillation using an inclusive
lepton trigger sample,\cite{cdfsltjqc} as shown in fig.
~\ref{fig:bbar}. The dilution of the soft lepton tag, as
measured using the $J/\psi K^+$ sample, is ${\cal D}$
= 63 $\pm$ 15 \%.
In the Jet Charge (JTQ) method a momentum weighted charge average of
particles
in a b-jet, $Q_{jet}$, is used to determine the charge of the b quark.
The event is considered as tagged when $|Q_{jet}|>0.2$.
The performance of the JTQ was also checked with the
analysis of the $\Delta m$ and dilution ${\cal D}$
using the inclusive lepton trigger sample (fig.
~\ref{fig:bbar}).
The dilution of the JTQ method, as measured with the
$J/\psi K^+$ sample, is ${\cal D}$ = 24 $\pm$ 7\%.
A summary of the performance of tagging algorithms, described by the
value of the dilution and the tagging efficiency,\footnote{
The tagging efficiency, $\epsilon$, is the fraction of
events that are tagged.} is presented in table~\ref{tab:eps}.
The Jet Charge and Soft Lepton tagging algorithms are described in more
detail in another CDF publication.\cite{cdfsltjqc}
\subsection{Same Side Tagging}\label{subsec:sst}
The Same Side Tagging (SST) technique relies on the correlation
between the flavor of the B hadron and the charge of a nearby
hadron produced either in the fragmentation process that
formed a B meson from a $b$ quark or from the decay of $B^{**}$ meson.
The charge correlations are the same in both cases:
a $B^0_d$ meson is associated with a positive particle.
The SST algorithm selects as a flavor tag, that particle which has
the minimum momentum component transverse to the momentum
sum of the B and the particle. The particle has to
be contained in an $\eta - \phi$
cone of 0.7 around the B momentum direction, have $P_t > 400~ MeV/c$,
and come from the primary vertex.
The performance of this method was calibrated by tagging $B \rightarrow
\ell D^{(*)}$ decays and observing the time dependence of the
$B^0_d \overline{B^0_d}$ oscillation,\cite{lepd}
as shown in fig. \ref{fig:bbar}.
In addition to the usual measurement of the frequency of the
oscillation $\Delta m_d$, the amplitude of the oscillation, $\cal{D}$,
called dilution,\footnote{The tagging dilution is defined as
${\cal{D}} = \frac{N_R - N_W}{N_R + N_W}$, where $N_W(N_R)$ are numbers of
wrong (right) tagging decisions. The observed asymmetry
${\cal A}_{CP}^{obs}={\cal D} {\cal A}_{CP}$.} was also determined.
\noindent
\section{Flavor Asymmetry in $\B0JKS$ Sample}
Following the conclusion of the Workshop, the CDF collaboration
updated the previously published analysis \cite{cdfsin2b} of the
$\B0JKS$ decays by
employing a combination of three tagging methods to the full sample
of $\sim$ 400 events. The SST and SLT algorithms were essentially
the same
as in those used in the $\ell D^{(*)}$\cite{lepd}
and inclusive lepton\cite{cdfsltjqc} analyses.
The JTQ algorithm
was modified from that in ref. 5
to increase the efficiency of identifying
low-$P_t$ jets. Each event can be tagged by one algorithm on the opposite
side and one on the same side. When both SLT and JTQ tags are
available, the tagger with higher dilution was selected
to avoid introduction of correlations. In addition, the dilution and
efficiencies for the opposite side tagging algorithms were determined
using a sample of $B \rightarrow J/\psi K^+$ decays, to match the
kinematic properties of the two samples.
\vspace{-0.5cm}
\begin{figure}[ht]
\centerline{
\epsfysize 6.5cm
\epsffile{ksh_vs_ct.eps}
\epsfysize 6.5cm
\epsffile{ctauk0s_asym_pr.eps}}
\caption{Results of CP asymmetry studies. Left: Time dependent
Same Side Tag applied to a sample of $\B0JKS$, where two muons are
reconstructed in the SVX detector. Right: Multiple tagging analysis results.
In addition to the time dependent information the plot displays
time-integrated asymmetry for non-SVX events. For comparison
the dashed line present results of the fit with $\Delta m_d$
left floating in the fit.
\label{fig:sin2beta}}
\end{figure}
Tagged events are simultaneously fitted for a combination of the three
tagging methods, using an unbinned likelihood fit with the value
of $\Delta m_d$ fixed to the world value.
The fitting also takes into account the remaining tag correlations.
The asymmetry values for the three tagging methods are shown in
fig. \ref{fig:sin2beta}. Those events
without proper time determination are presented separately as
a single point.
We measure $\sin(2\beta ) = 0.79 ^{+0.41}_{-0.44}$.
The curves shown in fig.~\ref{fig:sin2beta} present the results of the
fit.
\begin{table}[t]
\caption{Summary of the statistical power of the taggers, measured by
$\epsilon D^2$.\label{tab:eps}}
\vspace{0.2cm}
\begin{center}
\footnotesize
\begin{tabular}{|c|c|c|c|}
\hline
Tagger & \raisebox{0pt}[13pt][7pt]{Effective Dilution} &
\raisebox{0pt}[13pt][7pt]{Dilution} &{Efficiency}\\
& \raisebox{0pt}[13pt][7pt]{$\epsilon {\cal D}^2$} &
\raisebox{0pt}[13pt][7pt]{${\cal D}$} &{$\epsilon$}\\
\hline
SST & $2.1 \pm 0.5$ \% & 17 $\pm$ 4\% & 38$\pm$ 4\% \\
JTQ & $2.2 \pm 1.3$ \% & 24 $\pm$ 7\% & 40$\pm$ 4\% \\
SLT & $2.2 \pm 1.0$ \% & 63 $\pm$ 15\% & 6 $\pm$ 2\% \\
\hline
Multiple Tags & $6.3 \pm 1.7$ \% & & \\
\hline
\end{tabular}
\end{center}
\end{table}
\section{Conclusion}
Multiple tagging methods have been validated in the hadron collider
environment of the Tevatron. The statistical power of the taggers,
measured by the quantity $\epsilon {\cal D}^2$, was determined using data
sets accumulated by the CDF collaboration.
Using a sample of over $\sim 400$ events of fully reconstructed $\BJKS$
decays and multiple tags, we measured $\sin(2\beta ) =
0.79 ^{+0.41}_{-0.44}$.
This result can be translated into the frequentist confidence interval
of $0 < \sin(2\beta) <1$ at 93\% confidence level.
Next year, the Tevatron will begin a new collider run, delivering
an expected twenty-fold increase in data over the following two years.
With detector and trigger improvements, we expect to accumulate a sample of
$\sim$ 10,000 $\BJKS$ events,
allowing the uncertainty on $\sin(2\beta )$ to be reduced to 0.08.
All three tagging methods will be important tools in the study of
CP violation effects in the next run of the Tevatron.
\section*{Acknowledgments}
We thank the Fermilab staff and the technical staffs of the
participating institutions for their vital contributions. This work was
supported by the U.S. Department of Energy and National Science Foundation;
the Italian Istituto Nazionale di Fisica Nucleare; the Ministry of Education,
Science and Culture of Japan; the Natural Sciences and Engineering Research
Council of Canada; the National Science Council of the Republic of China;
the Swiss National Science Foundation; the A. P. Sloan Foundation; and the
Bundesministerium fuer Bildung und Forschung, Germany.
\section*{References}
|
\section{The Characteristic Photon Energy of Bursts}
The prompt emission of gamma-ray bursts is observed
predominately at several hundred keV. This is one
of the most interesting features of gamma-ray bursts,
since it is not easily explained by most theories of prompt
gamma-ray emission. In particular, the internal and
external shock theories predict a wide variation in
the characteristic gamma-ray energy both
during a burst and between bursts.
It is therefore imperative to know whether this is a
physical characteristic of bursts, or a consequence of
an instrumental effect.
In this article, we discuss the instrumental effects
that arise when observing gamma-ray bursts with
the Burst and Transient Source Experiment (BATSE) on the Compton
Gamma-Ray Observatory (CGRO).\cite{brainerd}
The primary points that
affect the observed
distribution function are the ability to correctly
define the characteristic energy of the gamma-ray burst and
the ability to trigger on a gamma-ray burst with
a characteristic energy outside of the trigger energy range.
These effects are discussed in \S 2 and \S 3 below. Models
of simple distributions for the characteristic energy are
given in \S 4, and their consistency with the observations
is discussed in \S 5. We find that the observed distribution
of characteristic energies cannot be explained through
instrumental effects alone. A model distribution with a
characteristic energy is fit to the observations to quantify
the type of physical theory required by the observations.
Our conclusion is that the existence
of approximately the same characteristic energy for all gamma-ray
burst spectra is a physical property of gamma-ray bursts that any
viable theory must explain.
\section{Defining E-Peak}
One can characterize the gamma-ray burst spectrum through the
E-peak value ($E_p$), the photon energy at which the $\nu F_{\nu}$ curve
has a maximum. The distribution of such values as measured by the
BATSE gamma-ray burst instrument is narrowly distributed.\cite{mallozzi}
The standard
method of modeling a gamma-ray spectrum is forward fitting. In this
method, a model photon spectrum is folded through a model of the
detector response matrix (DRM) to produce a count spectrum. The count
spectrum is then compared to the observed spectrum, with the best
fit found through $\chi^2$ minimization. The photon model used in
the analysis of the BATSE gamma-ray burst spectra is the gamma-ray burst
spectral form,\cite{band} which is a 4 parameter model that produces
good fits to the data. The data type used in deriving $E_p$ for
the BATSE data set is the MER data type, which covers the
spectrum with 16 channels.
To test the ability of BATSE to correctly determine the value of
$E_p$, we generated test burst count spectra with background counts added
for model spectra of a given $E_p$, and
then went through the procedure of subtracting background and deriving
the values of $E_p$ that provide the best fit to the test spectra.
We find that the forward fitting
method correctly defines the value of $E_p$ for
$20 \, \hbox{keV} < E_p < 2 \, \hbox{MeV}$. Once one is outside this range,
either no value of $E_p$ is found, or the value that is found is
at one of these two limits on $E_p$. Because few BATSE bursts have
$E_p$ at $\approx 2 \, \hbox{MeV}$, and no BATSE burst is at $\approx 20 \, \hbox{keV}$,
the misidentification of the value of $E_p$ for $E_p > 2 \, \hbox{MeV}$ and
$E_p < 20 \, \hbox{keV}$ is inconsequential.
\begin{figure}
\centering
\epsfig{figure=06_JBrainerd2_1.eps, width=4.0in}
\caption{Expected E-Peak distribution for bursts
with $C > C_{min}$ such that most bursts lie on the $-3/2$ portion
of the cumulative peak-flux distribution.
Here, $C$ is the burst
count rate in the energy range of the trigger and $C_{min}$ is
the limiting count rate for selecting bursts. The solid
curves are for triggering on an energy range of $50$--$300 \, \hbox{keV}$,
the dotted lines is for triggering on a range of $20$--$100 \, \hbox{keV}$,
and the dashed lines are for triggering on the range of $> 100 \, \hbox{keV}$.
The different curves are for different burst directions relative to Earth.
The spectral model used in these curves is a grb spectral form with
$\alpha = -1$ and $\beta = -3$.
}
\end{figure}
\begin{figure}
\centering
\epsfig{figure=06_JBrainerd2_2.eps, width=4.0in}
\caption{Expected E-Peak distribution for bursts with
$C > C_{min}$, with $C_{min}$ selected to place most bursts
on the $-0.8$ portion of the cumulative peak-flux
distribution. The remainder of the free parameters in the models is
as in Fig.~1.
}
\end{figure}
\section{BATSE Triggering}
The BATSE instrument triggers on a gamma-ray burst when the count rate
in the trigger channels exceed an average background count rate by
a preset value. Of the eight modules that comprise BATSE, at least two
must be above the background rate for the burst to trigger.
The triggering is done using four discriminator channels on
the timescales of $1.024 \, \hbox{s}$, $0.256 \, \hbox{s}$, and $0.064 \, \hbox{s}$.
For most bursts, the triggering is on
channels $2 + 3$, which contain counts in the $50 \, \hbox{keV}$ to
$300 \, \hbox{keV}$ energy range.
To model the trigger of a gamma-ray burst, we ran Monte Carlo simulations
of the count rates generated by a gamma-ray burst of a given normalization
and value of $E_p$. The randomness in this simulation was in the
orientation of the spacecraft relative to the gamma-ray burst and to
the center of the earth. Two aspects that affect the count rate
found for a burst are the orientation of all detectors to the burst, and
the component of the burst scattered into the detectors from Earth's
atmosphere. One aspect this simulation did not address was the
angular dependence of the background count rate. The purpose of the
analysis was to determine the average count rate in the second brightest
detector as a function of $E_p$. The result is used to
relate the photon rate of a burst to the count rate in the trigger
channels.
\begin{figure}
\centering
\epsfig{figure=06_JBrainerd2_3.eps, width=4.0in}
\caption{Best fit of power law E-peak distribution to
the BATSE E-peak distribution. The solid curve gives the
observed distribution of $E_p$. The dotted curve is the best
fit model, while the dashed curve is the best fit model renormalized
to match the maximum of the observed distribution.
}
\end{figure}
\begin{figure}
\centering
\epsfig{figure=06_JBrainerd2_4.eps, width=4.0in}
\caption{Best fit of log normal E-peak distribution to
the BATSE E-peak distribution. The solid curve is the observed
distribution, while the dotted curve is the best fit model.
}
\end{figure}
\section{Model E-Peak Distributions}
Based on the conversion of photon spectrum to count rate, one can
convert model distributions in $E_p$ and $N_p$, the photon rate
at $E_p$, into distributions in $E_p$ as functions of $C_{min}$,
the minimum count rate. In doing this, we assume that
the minimum count rate is high enough above the background that
threshold effects are unimportant. We are also ignoring the fact
that different detectors will have different background count rates.
We assume that the distribution of $N_p$ is a broken power law similar
to the peak-flux broken power law, with the power law having an index
of $-1.8$ below the break, and an index of $-5/2$ above the break.
The value at which power law index breaks in the $N_p$ distribution
is assumed to depend on $E_p$ as $E_p^{-2}$, which
is equivalent to the statement that the total energy per unit time emitted
as gamma-rays is independent of $E_p$ for bursts that fall on the break
of the peak-flux curve. Two models for the $E_p$ distribution are
used. The first is a power law, so that there is no characteristic
value of $E_p$ defined by the physics. The second is a log normal
curve with two power laws joined on to the wings of the distribution.
This second is used to fit the observed $E_p$ distribution to demonstrate the
extent to which the physical $E_p$ distribution must break to reproduce
the observations.
The model distributions for a power law $E_p$ distribution are given
in Figures 1 and 2. These curves are calculated for three different
trigger energy ranges: channels $1 + 2$ ($20 \, \hbox{keV}$--$100 \, \hbox{keV}$),
channels $2 + 3$ ($50 \, \hbox{keV}$--$300 \, \hbox{keV}$), and
channels $3 + 4$ ($> 100 \, \hbox{keV}$). The power law index for the $E_p$
distribution is set to $-1$. The difference between Figures 1 and 2 are
in the minimum count rate. The first has a minimum count rate such
that most bursts
are in the $-3/2$ portion of the cumulative peak-flux distribution,
while the second has a count rate such that most bursts are on
the $-0.8$ portion of the cumulative peak-flux distribution.
The effect of lowering the count rate limit is to broaden the
$E_p$ distribution.
\section{Fits of E-Peak Distributions}
The observed $E_p$ distribution was fit to the power law E-peak
distribution. Most bursts have low count rates, placing them
below the break in the peak-flux distribution, so the low count rate
model for the $E_p$ distribution is used.
We find that the best power law index for the $E_p$
distribution is $-0.94 \pm 0.11$, with $\chi^2 = 444.4$ for 8 degrees
of freedom. This is clearly a poor fit to the data. This model and the
observed $E_p$ distribution are given in Figure 3. The model distribution
is much broader than the observed distribution. This demonstrates that the
physical distribution of $E_p$ must have a characteristic value near
$200 \, \hbox{keV}$ to produce the observed distribution.
Just how narrow the physical distribution must be is demonstrated by the
fit of a log normal distribution to the observations. In this distribution,
the half maxima are a factor of 4 apart, and the power law tails have
indexes of $3.35$ and $-2.58$, so that the change in index exceeds 5.
The model fit has $\chi^2 = 7.93$ for 5 degrees of freedom, which is
an excellent fit to the data. The physical distribution therefore
must be very narrow to reproduce the observations.
This fit also suggests that above $1 \, \hbox{MeV}$, each decade in $E_p$
contains $< 10 \%$ of the bursts in the lower decade, and that
below $100\, \hbox{keV}$, each decade contains $< 1 \%$ of the bursts in the
higher decade.
\section{Discussion}
We have presented the results of our study of the BATSE instrumental
effects that affect the determination of the $E_p$ distribution.
Our results are as follows:
\begin{itemize}
\item Model fitting of spectra to gamma-ray burst spectra correctly
gives the value of $E_p$ for $20 \, \hbox{keV} < E_p < 2 \, \hbox{MeV}$.
\item The power law model curves for the $E_p$ distribution
for bursts that trigger on the $50$--$300 \, \hbox{keV}$ count energy range
have maxima at $\approx 100 \, \hbox{keV}$, and
half-maxima at $\approx 20 \, \hbox{keV}$ and $\approx 800 \, \hbox{MeV}$.
\item The power law model curves provide poor fits to the observations,
because the observations are at $\approx 10 \%$ of the peak values
at the energies where the model curves have their half-maxima.
\item Fitting a log normal $E_p$ distribution to the observations finds
that the best fit is very narrow, with power law wings that differ
by 5 in index.
\end{itemize}
These results demonstrate that the narrow $E_p$ distribution must have
a physical origin. Theories of the prompt gamma-ray emission must provide
an explanation for this property before they can be regarded as viable.
\section*{Acknowledgments}
This research was supported under the NASA grant NAG5-6746.
\section*{References}
|
\section{Introduction}
The selection of collective variables is an important problem
in the study of large amplitude collective motion. In the
constrained Hartree-Fock (CHF) or Hartree-Fock-Bogoliubov (CHFB)
calculations, the collective subspaces are generated by a small number
of one-body constraint (also called generalised cranking) operators
which are most
commonly taken to be multipoles that can be represented as
homogeneous polynomials in the coordinates, i.e., $r^L Y_{LK}$. In
realistic calculations of processes such as fission, the number of
coordinates needed to describe the full nuclear dynamics can easily become
larger than what can be dealt with in satisfactory manner, and a method to
determine the optimal combination needs to be devised.
Furthermore, there is no {\it a priori} reason to limit oneself to
multipole operators, and the cranking operators should be
determined by the nuclear collective dynamics itself, from amongst all
possible one-body operators.
In our past work, we have investigated a theory of adiabatic large
amplitude collective motion as a method to generate such self-consistent
collective subspaces (see Ref.~\cite{KWD91} and references
therein). The key ingredient of the method is the self-consistent
determination of the constraint operator, and as such it may provide an
answer to the selection question discussed above.
Using the local harmonic version of the theory \cite{KWD91},
the local harmonic approximation (LHA),
we have recently embarked on a study of the
properties of large amplitude collective motion in systems with
pairing. So far we have dealt with two simple models: The first one
is a semi-microscopic model of nucleons interacting through a pairing
force, coupled to a single harmonic variable \cite{NW98_1}. The second
\cite{NW98_2} is a fully microscopic $O(4)$ model which may be regarded as a
simplified version of the pairing-plus-quadrupole (P+Q) Hamiltonian studied
in this paper. It has turned out that the self-consistent collective
coordinate obtained by the LHA accounts quite well for the exact
dynamics of the models. We have also shown that the CHFB calculations
using the mass-quadrupole operator as the constraint operator can
result in incorrect results \cite{NW98_2}.
In this paper, which is an attempt to move towards realistic nuclear
problems, we will encounter some problems that need to be resolved first:
The main problem is the treatment of conservation laws.
In studies of large amplitude collective motion,
the mean-field states often do not have the symmetries of the Hamiltonian.
Then, the spurious (Nambu-Goldstone) degrees of freedom, such as
the translation and rotation of nucleus, inevitably appear,
which must be properly separated out from the other collective degrees of
freedom like nuclear shape change.
In the models discussed in Refs.~\cite{NW98_1,NW98_2},
the spurious degrees of freedom could
be removed by hand thanks to the simplicity of the models,
however in realistic nuclear problems, it is almost impossible to do so.
The second and more obvious problem
is the increase in the number of degrees of freedom.
Within the time-dependent Hartree-Bogoliubov (TDHB) approximation,
the dynamics of the models of Refs. \cite{NW98_1,NW98_2}
is described by only a modest number of degrees of freedom ($4\sim
12$). For realistic problems in heavy nuclei, on the other hand, we
need to deal with at least tens of thousands, and possibly
millions of degrees of freedom!
These two points, the effect of the conservation laws and the applications
to realistic problems,
are the main issues discussed in Sec.~\ref{sec: formalism}.
In the LHA, the collective path (or collective manifold for more than
one coordinate) is determined by solving the CHFB problem with a
cranking operator which is self-consistently determined by the local
random-phase approximation (RPA).
It is of course a well-known fact that the spurious modes decouple from the
physical modes in the RPA.
However this is true only when the system is at a minimum
of the mean-field potential.
Since the local RPA needs to be solved
at non-equilibrium points as well,
we need to study the properties of the LHA equations.
It will be shown that the symplectic version of the LHA
combined with an extended
adiabatic approximation properly takes account of conservation laws.
It will also be shown that this formalism naturally explains the role of the
chemical potential term $-\lambda N$ in the Hamiltonian
of the quasiparticle RPA.
Since the LHA procedure requires one to solve the local RPA at large number of
points along
the collective path (manifold), the large dimension of
the RPA matrix for realistic problems
leads to very heavy computational burdens.
Thus, instead of solving the local RPA,
we would like to preselect a small group
of operators, and
find the RPA modes as an optimal combination of these operators at each point of the
collective surface.
This means that we restrict
the RPA diagonalisation to this small space, rather than deal with
the full millions-by-millions RPA matrix \cite{DWK94}.
In this paper, we attempt to find a set of one-body operators
which approximate well the self-consistent cranking operator.
The pairing-plus-quadrupole (P+Q) Hamiltonian is adopted for the analysis
because the RPA calculation
is relatively easy due to the separability of the force (some of the results
for the Sm isotopes have already been reported in Ref.~\cite{NWD99}).
Since the P+Q model is known to be able to describe collective
phenomena involving both pairing and quadrupole degrees of freedom
\cite{Bel59,KS60,Bes61,KS63,BK_I,BK_II,BK_III,BK_IV,BK_V},
we expect that a similar choice for the set of
one-body operators should work for other realistic Hamiltonians as well.
The P+Q model is probably one of the most simple and successful
nuclear Hamiltonians which allows us to discuss realistic problems
involving pairing and quadrupole degrees of freedom. Baranger and
Kumar analysed in great detail the (adiabatic) collective motion in
the P+Q model assuming that the collective variables are the mass
quadrupole operators \cite{BK_I,BK_II,BK_III,BK_IV,BK_V}.
Thus, they reduced the large number
of two-quasiparticle (2qp) degrees of freedom (for this model of the order of
a thousand) to only two ``collective'' coordinates, $\beta$ and
$\gamma$. However, our previous study of the $O(4)$ model\cite{NW98_2}
suggests that even for such simple Hamiltonians, the self-consistent
collective coordinates are often not as simple as one might think.
The P+Q model (with an additional quadrupole pairing interaction)
has also been studied by Kishimoto and Tamura \cite{KT72,KT76}
using boson expansion techniques. They found that
the coupling to (non-collective) 2qp excitations is
very important, which also suggests that the collective coordinates are
not just the mass-quadrupole ones.
In this paper we show that the
normal-mode coordinate of the random-phase approximation is
quite different from the mass-quadrupole operator,
and how the structure of the self-consistent collective coordinates
changes as the nucleus is deformed.
In Sec.~\ref{sec: formalism}, we give an account of the general
theoretical arguments.
The problem of the spurious modes and the extension of the RPA formalism
are discussed in Sec.~\ref{sec: RLHA} and \ref{sec: SLHA}.
In Sec.~\ref{sec: PLHA}, the method of projection of the RPA equation
onto a set of one-body operators is described.
Numerical calculations for the RPA and the projected RPA for the P+Q model
are given in Sec.~\ref{sec: P_plus_Q}
and finally we give the summary and outlook in Sec.~\ref{sec: conclusions}.
\section{Local harmonic formulations and conservation laws}
\label{sec: formalism}
In this section, we use a summation convention where
the repeated appearance of the same symbol
for upper and lower index
indicates a sum over this symbol for all allowed values.
We also use the convention that a comma in a lower index
indicates the derivative with respect to the coordinate,
i.e., $F_{,\alpha} = \partial F/\partial \xi^\alpha$.
\subsection{The TDHFB theory and classical constants of motion}
Utilising the TDHFB theory
we transcribe the original quantum Hamiltonian
into the classical Hamiltonian.
We follow Ref.~\cite{BR86} to define the canonical variables.
We denote a time-dependent generalised Slater determinant by
$\ket{\Psi(t)}$ which is characterised by
the (quasi-)density and pair matrices,
\begin{equation}
\bar{\rho}_{ij} \equiv
\bra{\Psi(t)} a_j^\dagger a_i \ket{\Psi(t)} ,\quad\quad
\bar{\kappa}_{ij} \equiv
\bra{\Psi(t)} a_j a_i \ket{\Psi(t)} ,
\end{equation}
where $\ket{\Psi(t)}$ is assumed to be normalised at all times
($\inproduct{\Psi(t)}{\Psi(t)}=1$) and
the quasiparticle operators $(a_i, a_i^\dagger)$ are
defined with respect to a reference state $\ket{\Psi_0}$.
Based on a classical Holstein-Primakoff mapping \cite{HP40,KM90},
\begin{equation}
\label{H_P_mapping}
\bar{\rho}_{ij}
= \left[ \beta \beta^\dagger \right]_{ij} , \quad\quad
\label{quasi_pairing_tensor}
\bar{\kappa}_{ij}
= \left[ \beta (1 - \beta^\dagger \beta)^{1/2} \right]_{ij} ,
\end{equation}
the density and pair matrices are now mapped onto complex canonical variables
$\beta_{ij}$.
As usual real canonical variables $\xi$ and $\pi$ are introduced as
$\beta = (\xi + i \pi) / \sqrt{2} $.
In this canonical parametrisation,
the reference state $\ket{\Psi_0}$ corresponds to
the origin of the phase space, where $\xi=\pi=0$.
We shall specifically look at
one-body hermitian operators $R$ and anti-hermitian operators $S$,
\begin{eqnarray}
R &=& R_0 + \sum_{i > j} R(ij)
(a_i^\dagger a_j^\dagger + a_j a_i )
+ \sum_{ij} R'(ij) a_i^\dagger a_j ,\\
S &=& \sum_{i > j} S(ij)
(a_i^\dagger a_j^\dagger - a_j a_i )
+ \sum_{ij} S'(ij) a_i^\dagger a_j ,
\end{eqnarray}
where $R_0=\bra{\Psi_0} R \ket{\Psi_0}$ and
all the matrix elements, $R(ij)$, $R'(ij)$, $S(ij)$ and $S'(ij)$,
are assumed to be real.
Here $R'(ij)$ ($S'(ij)$) are symmetric (anti-symmetric) with respect to
permutation of the indices $i$ and $j$.
The corresponding classical representations,
${\cal R}(\xi,\pi)$ and ${\cal S}(\xi,\pi)$, can be found by
replacing the fermion pair operators $a_j^\dagger a_i$ and $a_j a_i$
by the density and pair matrices
$\bar\rho_{ij}$ and $\bar\kappa_{ij}$ of Eq.~(\ref{H_P_mapping}).
Expanding $\bar\rho$ and $\bar\kappa$ with respect to $\pi$,
we find that ${\cal R}(\xi,\pi)$ consists of even-order terms in $\pi$
while ${\cal S}(\xi,\pi)$ contains odd-order terms only,
\begin{eqnarray}
\label{classical_R}
{\cal R} (\xi, \pi) &=&
{\cal R}^{(0)}(\xi)
+ \frac{1}{2} {\cal R}^{(2)\alpha\beta}(\xi) \pi_\alpha \pi_\beta
+ {\cal O}(\pi^4) , \\
\label{classical_S}
{\cal S} (\xi, \pi) &=&
{\cal S}^{(1)\alpha}(\xi) \pi_\alpha + {\cal O}(\pi^3) .
\end{eqnarray}
Starting from these equations,
pairs of two-quasiparticle (2qp) indices $(ij,kl,\cdots)$
are denoted collectively by a single Greek index ($\alpha$, $\beta$, $\cdots$).
The 2qp matrix elements with respect to the reference state,
$R(ij)$ and $S(ij)$ which are now denoted as $R(\alpha)$ and $S(\alpha)$
respectively,
are related to the derivatives of the classical representations,
\begin{eqnarray}
\label{2qp_R}
R(\alpha) &=&
\frac{1}{\sqrt{2}}
\left.\frac{\partial {\cal R}}{\partial \xi^\alpha}\right|_{\xi=\pi=0}
=\frac{1}{\sqrt{2}} \left. {\cal R}^{(0)}_{,\alpha}\right|_{\xi=0} ,\\
\label{2qp_S}
S(\alpha) &=&
\frac{i}{\sqrt{2}}
\left.\frac{\partial {\cal S}}{\partial \pi_\alpha}\right|_{\xi=\pi=0}
=\frac{i}{\sqrt{2}} \left. {\cal S}^{(1)\alpha}\right|_{\xi=0} .
\end{eqnarray}
The classical Hamiltonian is given by
\begin{equation}
{\cal H} (\xi, \pi) = \bra{\Psi(t)} H \ket{\Psi(t)} ,
\end{equation}
where the right hand side is calculated in terms of
the density and pair matrices (\ref{H_P_mapping}).
The expansion of ${\cal H}(\xi,\pi)$ in powers of $\pi$ defines the potential
$V(\xi) = {\cal H} (\pi=0)$ (zeroth order)
and the mass parameter $B^{\alpha\beta}(\xi)$ (second order).
\begin{eqnarray}
\label{adiabatic_Hamiltonian}
{\cal H} &=& V(\xi) + \frac{1}{2} B^{\alpha\beta} \pi_\alpha \pi_\beta
+ {\cal O}(\pi^4) ,\\
B^{\alpha\beta}(\xi) &=&
\left. \frac{\partial^2 {\cal H}}{\partial\pi_\alpha \partial\pi_\beta}
\right|_{\pi=0} .
\end{eqnarray}
The terms of order of $\pi^4$ and higher in the Hamiltonian
are neglected in the adiabatic time-dependent mean-field theory.
The tensor $B_{\alpha \beta}$, which is defined as the inverse of
$B^{\alpha \beta}$
($B^{\alpha \gamma} B_{\gamma \beta} = \delta^\alpha_\beta$),
plays the role of metric tensor in the Riemannian formulation of
the local harmonic approximation (LHA) as discussed in Sec.~\ref{sec: RLHA}.
It is well-known that the TDHFB equation preserves the symmetries
of the quantum Hamiltonian:
If the one-body operator $P$ commutes with the Hamiltonian,
$[P,H]=0$, the classical representation of $P$,
${\cal P}(\xi,\pi)=\bra{\Psi(t)} P \ket{\Psi(t)}$, is
a classical constant of motion.
Consequently, the Poisson bracket between ${\cal P}$ and ${\cal H}$ must vanish,
\begin{equation}
\{ {\cal P}, {\cal H} \}_{\rm PB} = 0 ,
\label{eq:2.12}
\end{equation}
which can be used to show that
\begin{eqnarray}
\label{TDHFB_eq_1}
{\cal P}^{(1)\alpha} V_{,\alpha} = 0 ,\\
\label{TDHFB_eq_2}
{\cal P}^{(0)}_{,\alpha} B^{\alpha\beta}
- {\cal P}^{(2)\alpha\beta} V_{,\alpha} = 0 ,
\end{eqnarray}
which are the terms of zeroth and first order in $\pi$ in Eq.~(\ref{eq:2.12}),
respectively.
Here, ${\cal P}^{(0)}$, ${\cal P}^{(1)\alpha}$ and
${\cal P}^{(2)\alpha\beta}$ are defined,
according to Eqs.~(\ref{classical_R}) and (\ref{classical_S}),
as terms of the zeroth, first and second order in $\pi$, respectively.
The equations (\ref{TDHFB_eq_1}) and (\ref{TDHFB_eq_2})
hold at arbitrary points in configuration space.
\subsection{Riemannian version of the local harmonic approximation (LHA)}
\label{sec: RLHA}
The adiabatic approach to large amplitude collective motion is
based on a search for the collective (and non-collective) coordinates $q^\mu$
by performing a point transformation
of the original coordinates $\xi^\alpha$,
\begin{equation}
\label{point_transf_1}
q^\mu = f^\mu(\xi) , \quad\quad
\xi^\alpha = g^\alpha(q) \quad\quad (\mu,\alpha=1,\cdots,n).
\end{equation}
The conjugate momenta are given by
\begin{equation}
\label{point_transf_2}
p_\mu = g^\alpha_{,\mu} \pi_\alpha , \quad\quad
\pi_\alpha = f^\mu_{,\alpha} p_\mu .
\end{equation}
The adiabatic Hamiltonian (\ref{adiabatic_Hamiltonian}) is transformed into
\begin{eqnarray}
\bar{{\cal H}} &=& \bar{V}(q) + \frac{1}{2} \bar{B}^{\mu\nu} p_\mu p_\nu
+ {\cal O}(p^4) ,\\
\bar{V}(q) &=& V(g(q)) , \quad\quad
\bar{B}^{\mu\nu} = f^\mu_{,\alpha} B^{\alpha\beta} f^\nu_{,\beta} .
\end{eqnarray}
The point transformation on the adiabatic Hamiltonian gives the same result
as the one obtained by first performing this transformation on the exact
Hamiltonian and then taking the adiabatic limit.
This is not true for a general canonical transformation,
specifically for the extended adiabatic transformation
to be discussed in the next section
We divide the new coordinate set $q^\mu$ into three subsets,
the collective coordinates $q^1$,
the spurious coordinates $q^I$, $I=2,\cdots,M+1$, and
the non-collective coordinates $q^a$, $a=M+2,\cdots,n$.
For simplicity, we assume here a single collective coordinate.
The spurious coordinates are related to the symmetry breaking in the
mean field (Nambu-Goldstone modes).
The local harmonic approximation consists of two sets of equations,
the force condition and the local RPA,
which define the collective manifold in configuration space
(see Ref.~\cite{KWD91} for a complete description).
When spurious modes are present,
the force condition is modified slightly,
\begin{equation}
\label{force_condition}
V_{,\alpha} = \lambda_1 f^1_{,\alpha} + \lambda_I f^I_{,\alpha} ,
\end{equation}
where the second term in the right-hand side is the additional term.
The local RPA equation can be formulated in different ways \cite{KWD91},
and in this section we discuss the case of the Riemannian formulation,
based on
\begin{eqnarray}
\label{RLHA_1}
V_{;\alpha\gamma} B^{\gamma\beta} f^1_{,\beta}
&=& (\hbar\Omega)^2 f^1_{,\alpha}, \\
\label{RLHA_2}
g^\alpha_{,1} V_{;\alpha\gamma} B^{\gamma\beta}
&=& (\hbar\Omega)^2 g^\beta_{,1}.
\end{eqnarray}
Here the covariant derivative $V_{;\alpha\gamma}$ is defined by
\begin{equation}
V_{;\alpha\beta} \equiv V_{,\alpha\beta}
- \Gamma^\gamma_{\alpha\beta} V_{,\gamma} ,
\end{equation}
where the affine connection $\Gamma$ is defined with the help of
metric tensor $B_{\alpha\beta}$ as
\begin{equation}
\label{affine_1}
\Gamma^\alpha_{\beta\gamma} = \frac{1}{2}
B^{\alpha\delta} \left( B_{\delta\beta, \gamma}
+ B_{\delta\gamma , \beta}
- B_{\beta\gamma , \delta} \right) .
\end{equation}
The collective path will be
determined by solving Eqs.~(\ref{force_condition}) and (\ref{RLHA_1})
selfconsistently.
It consists in finding a series of points where the local RPA eigenvector
$f^1_{,\alpha}$ satisfies the force condition at the same time.
The question at present is whether the solution of Eqs.~(\ref{RLHA_1}) and
(\ref{RLHA_2}) is orthogonal to the spurious modes.
The spurious coordinates are normally linked to the one-body
operators $P_I$ which commute with the Hamiltonian.
It is convenient to discuss separately the case
in which the symmetry operators $P_I$
are hermitian with real matrix elements
and the case in which $P_I$ are anti-hermitian with
real matrix elements.
First, let us discuss the latter case.
As is seen in Eq.~(\ref{classical_S}), neglecting the higher-order terms
in $\pi$,
the classical representation of the symmetry operator
is a classical momentum variable $p_I$ that is a constant of motion,
\begin{equation}
p_I = \bra{\Psi} P_I \ket{\Psi}
= {\cal P}_I^{(1)\alpha} \pi_\alpha + {\cal O}(\pi^3) .
\end{equation}
From this, we can immediately see that
$g^\alpha_{,I} = {\cal P}_I^{(1)\alpha}$.
Differentiating Eq.~(\ref{TDHFB_eq_1}) with respect to $\xi^\beta$,
keeping in mind that
${\cal P}^{(1)\alpha}$ must be proportional to $g^\alpha_{,I}$,
we find
\begin{equation}
\label{TDHFB_eq_3}
V_{,\alpha\beta} g^\alpha_{,I}
+ V_{,\alpha} g^\alpha_{,I\mu} f^\mu_{,\beta} = 0 .
\end{equation}
Using an identity for the coordinate transformation \cite{KWD91},
\begin{equation}
g^\alpha_{,\mu\nu}
+ \Gamma^\alpha_{\beta\gamma} g^\beta_{,\mu} g^\gamma_{,\nu}
- \bar{\Gamma}^\lambda_{\mu\nu} g^\alpha_{,\lambda} = 0 ,
\end{equation}
this leads to
\begin{equation}
\label{RLHA_NG_1}
g^\alpha_{,I} V_{;\alpha\gamma} B^{\gamma\beta} =
- \bar{V}_{,\lambda} \bar{\Gamma}^\lambda_{I\mu}
\bar{B}^{\mu\nu} g^\beta_{,\nu} ,
\end{equation}
where $\bar{\Gamma}$ is defined in the same way as $\Gamma$,
(\ref{affine_1}), with obvious change of $B$ to $\bar{B}$.
At the equilibrium $V_{,\alpha} = \bar{V}_{,\mu} = 0$,
and since the right-hand side of Eq.~(\ref{RLHA_NG_1}) vanishes,
the spurious modes are indeed
the zero-energy solutions of the RPA equation.
Thus, all finite-energy solutions are automatically
orthogonal to the spurious modes.
However, at non-equilibrium points, the $g^\alpha_{,I}$ are no longer
RPA eigenmodes and there is no guarantee that the collective mode
$g^\alpha_{,1}$ obtained from the local RPA equation (\ref{RLHA_2})
is free from spurious admixture\footnote{
If $\bar{B}^{\mu\nu}$ and $\bar{\Gamma}^1_{\mu\nu}$
are block-diagonal for the spurious and the collective spaces
on the collective path,
the spurious solutions correspond to finite energy eigenvalues but
are orthogonal to the collective mode.
This is the case for HF problem in $^{28}$Si as discussed in Refs.~\cite{DWK94,WDK91,WKD92}.}.
In Ref.~\cite{NW98_1}, the local RPA equation was modified to
insure that $f^I_{,\alpha} g^\alpha_{,1} = g^\alpha_{,I} f^1_{,\alpha} =0$.
However, in this formulation, we need to calculate both $f^I_{,\alpha}$
and $g^\alpha_{,I}$, which means that we have to solve the
Thouless-Valatin equations for the spurious modes \cite{TV62,MW69}.
In the second case, where the symmetry operator $P_I$ is hermitian,
we have coordinates $q^I$ which are the constants of motion:
\begin{equation}
\label{q_point_transf}
q^I = \bra{\Psi} P_I \ket{\Psi}
= {\cal P}_I^{(0)}(\xi)
+ \frac{1}{2} {\cal P}_I^{(2)\alpha\beta} \pi_\alpha \pi_\beta
+ {\cal O}(\pi^4) .
\end{equation}
If we limit ourselves to point transformations, we find
${\cal P}_I^{(0)}(\xi) = f^I (\xi)$ and we should ignore all terms depending on $\pi$.
At equilibrium Eq.~(\ref{TDHFB_eq_2}) implies that
the spurious modes $f^I_{,\alpha}$ correspond to
zero eigenmode of the mass matrix $B^{\alpha\beta} f^I_{,\beta} = 0$.
Therefore we have a zero-energy mode,
\begin{equation}
\label{RLHA_NG_2}
V_{;\beta\gamma} B^{\gamma\alpha} f^I_{,\alpha} = 0 .
\end{equation}
Away from equilibrium, however, the $f^I_{,\alpha}$ are no longer RPA eigenmodes, due to the second term in (\ref{TDHFB_eq_2})
which is related to the terms we have neglected in Eq.~(\ref{q_point_transf}).
In summary, the Riemannian LHA can automatically separate out the spurious
modes only at equilibrium $V_{,\alpha}=0$.
Since the local RPA at equilibrium is nothing but the conventional RPA,
the decoupling of the spurious modes is well-known.
However, at non-equilibrium points, the spurious modes can in general mix
with the collective one.
This undesirable mixing disappears in the symplectic LHA with an extended
adiabatic treatment described in the next section.
\subsection{Symplectic LHA and extended adiabatic transformation}
\label{sec: SLHA}
The symplectic version of the LHA \cite{KWD91}
is formulated with an affine connection
\begin{equation}
\widetilde{\Gamma}^\alpha_{\beta\gamma}
\equiv g^\alpha_{,\mu} f^\mu_{,\beta\gamma} ,
\end{equation}
which can be written in the usual form (\ref{affine_1}) by replacing
the metric tensor $B_{\alpha\beta}$ by
\begin{equation}
K_{\alpha\beta} \equiv \sum_\mu f^\mu_{,\alpha} f^\mu_{,\beta} .
\end{equation}
Since this metric depends on the final coordinates $q^\mu$,
we cannot define it from the beginning in contrast to the
Riemannian metric $B_{\alpha\beta}$.
In this formulation,
the configuration space is
assumed to be flat and diagonal ($K_{\mu\nu} = \delta_{\mu\nu}$)
in the final coordinate system $q^\mu$.
The covariant derivative is defined by
\begin{eqnarray}
\label{SLHA_cov_dev}
\widetilde{V}_{;\alpha\beta} &=&
V_{,\alpha\beta}
- \widetilde{\Gamma}^\gamma_{\alpha\beta} V_{,\gamma} \nonumber \\
&=& V_{,\alpha\beta} - f^\mu_{,\alpha\beta} \bar{V}_{,\mu} .
\end{eqnarray}
For the case that the momenta $p_I$ are constants of motion,
we can prove that the spurious modes are the zero-energy solutions
of the symplectic local RPA equation.
Indeed, differentiating the chain relation
$g^\alpha_{,\mu} f^\mu_{,\beta} = \delta^\alpha_\beta$
with respect to $q^\nu$, we obtain
\begin{equation}
g^\alpha_{,\mu\nu} f^\mu_{,\beta}
= - g^\alpha_{,\mu} g^\gamma_{,\nu} f^\mu_{,\beta\gamma} .
\end{equation}
Utilising this equation and the definition of the covariant derivatives
(\ref{SLHA_cov_dev}),
Eq.~(\ref{TDHFB_eq_3}) can be rewritten as
\begin{equation}
\widetilde{V}_{;\alpha\beta} g^\alpha_{,I} = 0 .
\end{equation}
This is a consequence of the TDHFB equation of motion.
Therefore the spurious modes are zero-energy solutions of the
symplectic LHA equations, anywhere in the configuration space.
In the case that the coordinates $q^I$ are constants of motion,
we need to lift the restriction to point transformations,
based on the power expansion with respect to $\pi$ keeping the
adiabatic assumption.
Instead of Eqs.~(\ref{point_transf_1}) and (\ref{point_transf_2}),
we generalise these to
\begin{eqnarray}
\label{EAT_0}
q^\mu &=& f^\mu(\xi)
+ \frac{1}{2} f^{(1)\mu\alpha\beta}(\xi) \pi_\alpha \pi_\beta
+ {\cal O}(\pi^4) , \\
\label{EAT_0b}
\xi^\alpha &=& g^\alpha(q)
+ \frac{1}{2} g^{(1)\alpha\mu\nu}(q) p_\mu p_\nu
+ {\cal O}(p^4) ,
\end{eqnarray}
and
\begin{eqnarray}
p_\mu &=& g^\alpha_{,\mu} \pi_\alpha + {\cal O}(\pi^3) ,\\
\pi_\alpha &=& f^\mu_{,\alpha} p_\mu + {\cal O}(p^3) ,
\label{eq:2.37}
\end{eqnarray}
where the terms cubic in momenta do not play a role in the modification
of the theory.
The original Hamiltonian (\ref{adiabatic_Hamiltonian}) is transformed to,
up to second order in $p$ only,
\begin{eqnarray}
\bar{H}(q,p) &=& \bar{V}(q) + \frac{1}{2} \bar{B}^{\mu\nu} p_\mu p_\nu ,\\
\bar{B}^{\mu\nu} &=& f^\mu_{,\alpha} B^{\alpha\beta} f^\nu_{,\beta}
+ V_{,\gamma} g^{(1)\gamma\mu\nu} .
\end{eqnarray}
Substituting Eq.~(\ref{EAT_0}) in Eq.~(\ref{EAT_0b}), using Eq.~(\ref{eq:2.37}),
we find the relation
\begin{equation}
\label{EAT_1}
g^{(1)\alpha\mu\nu} g^\beta_{,\mu} g^\gamma_{,\nu} =
-f^{(1)\lambda\beta\gamma} g^\alpha_{,\lambda} .
\end{equation}
From the canonicity condition $\{ q^\mu, q^\nu \}_{\rm PB} = 0$,
we also find
\begin{equation}
\label{EAT_2}
f^\mu_{,\alpha} f^{(1)\nu\alpha\beta}
= f^\nu_{,\alpha} f^{(1)\mu\alpha\beta} .
\end{equation}
The major difference between the use of the extended adiabatic
transformation and a point transformation is the modification
of mass parameter,
\begin{eqnarray}
\label{modified_mass}
\widetilde{B}^{\alpha\beta} &\equiv&
g^\alpha_{,\mu} \bar{B}^{\mu\nu} g^\beta_{,\nu} \nonumber \\
&=& B^{\alpha\beta} - \bar{V}_{,\mu} f^{(1)\mu\alpha\beta} .
\end{eqnarray}
Here we have used the relation (\ref{EAT_1}) to obtain the last equation.
The local RPA equations have the same forms as the Riemannian RPA equations,
(\ref{RLHA_1}) and (\ref{RLHA_2}),
after replacing $V_{;\alpha\beta}$ by $\widetilde{V}_{;\alpha\beta}$ and
$B^{\alpha\beta}$ by $\widetilde{B}^{\alpha\beta}$.
Let us return to the spurious modes.
For hermitian constants of motion, i.e., coordinates $q^I$
as in Eq.~(\ref{q_point_transf}),
we can identify $f^I(\xi) = {\cal P}^{(0)}_I(\xi)$ and
$f^{(1)I\alpha\beta} = {\cal P}^{(2)\alpha\beta}_I$
for the extended adiabatic transformation.
Utilising the TDHFB equation (\ref{TDHFB_eq_2})
and the canonicity condition (\ref{EAT_2}), we find
\begin{eqnarray}
\label{EAT_3}
\widetilde{B}^{\alpha\beta} f^I_{,\beta} &=&
B^{\alpha\beta} f^I_{,\beta}
- f^{(1)\mu\alpha\beta} \bar{V}_{,\mu} f^I_{,\beta} \nonumber \\
&=& B^{\alpha\beta} f^I_{,\beta}
- f^{(1)I\alpha\beta} \bar{V}_{,\mu} f^\mu_{,\beta} \nonumber \\
&=& B^{\alpha\beta} f^I_{,\beta}
- f^{(1)I\alpha\beta} V_{,\beta} \nonumber \\
&=& 0 .
\end{eqnarray}
Therefore, anywhere in the space, it is guaranteed that
the spurious modes are zero-energy solutions of
the symplectic LHA with the extended adiabatic transformation.
It is illustrative to discuss the difference between
the Hartree-Fock (HF) and Hartree-Fock-Bogoliubov (HFB) approaches.
The HF state is at an equilibrium point of the potential, $V_{,\alpha}=0$,
where the force condition (\ref{force_condition}) is satisfied
with $\lambda_1=\lambda_I=0$.
We also see
$\widetilde{V}_{;\alpha\beta} = V_{;\alpha\beta} = V_{,\alpha\beta}$
and $\widetilde{B}^{\alpha\beta} = B^{\alpha\beta}$.
Thus, any version of the LHA is equivalent to the conventional RPA.
On the other hand, the HFB state is not a real equilibrium.
We have a spurious coordinate associated with the particle number
$q^N = \bra{\Psi} N \ket{\Psi}$
and the gradient of the potential has a non-zero component
along the direction of $q^N$, $\bar{V}_{,N} \neq 0$
(the other components of the gradient are all zero).
The force condition is again satisfied with $\lambda_1=0$ and
$\lambda_N=\bar{V}_{,N}$, but neither
$\widetilde{V}_{;\alpha\beta} = V_{;\alpha\beta}$ nor
$\widetilde{B}^{\alpha\beta} = B^{\alpha\beta}$ hold any more.
Instead, we have
\begin{eqnarray}
\label{EAT_HFB}
\widetilde{B}^{\alpha\beta} &=& B^{\alpha\beta}
- \bar{V}_{,N} f^{(1)N\alpha\beta} ,\\
\label{SLHA_HFB}
\widetilde{V}_{;\alpha\beta} &=& V_{,\alpha\beta}
- \bar{V}_{,N} f^N_{,\alpha\beta} .
\end{eqnarray}
Since $\bar{V}_{,N}$ is nothing but the chemical potential $\lambda_N$
at the (constrained) HFB minimum,
the symplectic version of the LHA with the extended adiabatic approximation
is equivalent to the conventional quasiparticle RPA
with the constrained Hamiltonian, $H' = H - \lambda_N N$.
This guarantees the separation of zero-energy spurious modes,
while this is not the case for the Riemannian version of the LHA.
This symplectic formulation also illuminates
why the quasiparticle RPA at the HFB state
should be performed using the Hamiltonian $H'$, not the original $H$.
As we have seen above, the symplectic formulation of LHA with the extended
adiabatic transformation has an advantage over its Riemannian version.
Unfortunately we do not know a general method to calculate
$f^{(1)\mu\alpha\beta}$.
It may be possible to calculate this quantity
if the collective coordinate is explicitly given by
a combination of elementary one-body operators (see the discussion in Sec.~\ref{sec: conclusions}).
\subsection{Projected LHA and the spurious modes}
\label{sec: PLHA}
One of the main obstacles in applying the LHA techniques described in previous
sections to realistic nuclear problems is the fact that we need to diagonalise
an RPA matrix of large dimensionality at each point of the collective path.
In this section, we describe a method to approximate the RPA eigenvector
by taking linear combinations of preselected one-body operators,
generalising the method discussed in Ref. \cite{DWK94}.
First we select a small number of one-body operators,
$F^{(k)}$, $k=1,\cdots,n'$,
assuming that the collective coordinate
$q^1$
can be approximated locally as a
linear combination of these operators,
\begin{equation}
q^1 \approx \bra{\Psi} \hat{f} \ket{\Psi} , \quad
\hat{f} = \sum_{k=1}^{n'} c_k F^{(k)} .
\end{equation}
This means that the RPA eigenmodes $f_{,\alpha}$ is projected onto
the subspace spanned by $\{ {\cal F}^{(k)}_{,\alpha} \}$,
\begin{equation}
f_{,\alpha} \approx \bar{f}_{,\alpha} = \sum_{k=1}^{n'} c_k {\cal F}^{(k)}_{,\alpha} ,
\end{equation}
where ${\cal F}^{(k)}$ are the classical representations
(expectation values) of $F^{(k)}$.
In order to determine the coefficients $c_k$ in the linear combinations,
the local RPA equation is also projected onto the subspace
$\{ {\cal F}^{(k)}_{,\alpha} \}$:
\begin{equation}
\label{projected_LHA}
{\bf M}^{kl} c_l = (\hbar\bar{\Omega})^2 {\bf N}^{kl} c_l ,\\
\end{equation}
where $\hbar\bar{\Omega}$ is a frequency of the projected RPA and
\begin{eqnarray}
\label{projected_LHA_M}
{\bf M}^{kl} &=& {\cal F}^{(k)}_{,\alpha} B^{\alpha\beta}
V_{;\beta\gamma} B^{\gamma\delta} {\cal F}^{(l)}_{,\delta} ,\\
\label{projected_LHA_N}
{\bf N}^{kl} &=& {\cal F}^{(k)}_{,\alpha} B^{\alpha\beta} {\cal F}^{(l)}_{,\beta} .
\end{eqnarray}
Normally, in order to obtain the RPA eigenmodes and frequencies,
we need to diagonalise the RPA matrix
$B^{\alpha\gamma} V_{;\gamma\beta}$
whose dimension is equal to the number of active 2qp degrees of freedom.
The dimension of matrices ${\bf M}^{kl}$ and
${\bf N}^{kl}$ is equal to the number of selected one-body
operators $\{F^{(k)}\}$.
Therefore, if we can approximate the RPA
eigenvectors by using a small number of operators, it will
significantly reduce the computational task.
The equation for the force condition is also projected the same way,
\begin{equation}
\label{projected_force_condition_1}
{\bf V}^{k} = \lambda_1 {\bf N}^{kl} c_l ,\\
\end{equation}
with
\begin{equation}
\label{projected_LHA_V}
{\bf V}^{k} = {\cal F}^{(k)}_{,\alpha} B^{\alpha\beta} V_{,\beta} .
\end{equation}
Equations (\ref{projected_LHA}) and
(\ref{projected_force_condition_1}) must be solved self-consistently
in order to determine a collective path.
In cases where spurious modes exist,
the spurious components should be removed from each operator as
\begin{equation}
\label{PLHA_op}
\bar{\cal F}^{(k)}_{,\alpha} = {\cal F}^{(k)}_{,\alpha}
- {\cal F}^{(k)}_{,\beta} g^\beta_{,I} f^I_{,\alpha} ,
\end{equation}
and the force condition (\ref{projected_force_condition_1}) should be also
modified by adding the terms $\lambda_I {\bf W}^{kI}$ in the right-hand
side with
${\bf W}^{kI} = \bar{\cal F}^{(k)}_{,\alpha} B^{\alpha\beta} f^I_{,\beta}$.
The removal of spurious modes in Eq.~(\ref{PLHA_op}) is a tedious task
because we need to obtain both $f^I_{,\alpha}$ and $g^\alpha_{,I}$
by solving the Thouless-Valatin equation.
However, in case of hermitian constants of motion $q^I$,
adopting the extended adiabatic approximation in the previous section,
we show that the explicit removal as in Eq.~(\ref{PLHA_op}) is unnecessary.
Using the modified mass parameter (\ref{modified_mass}),
we have $\widetilde{B}^{\alpha\beta} f^I_{,\beta}=0$, Eq.~(\ref{EAT_3})
showing that
the $f^I_{,\alpha}$ correspond to the zero eigenmodes of mass matrix
(\ref{EAT_3}).
Therefore, we find
\begin{eqnarray}
{\bf M}^{kl} &=& {\cal F}^{(k)}_{,\alpha} B^{\alpha\beta}
V_{;\beta\gamma} B^{\gamma\delta} {\cal F}^{(l)}_{,\delta}
= \bar{\cal F}^{(k)}_{,\alpha} B^{\alpha\beta}
V_{;\beta\gamma} B^{\gamma\delta} \bar{\cal F}^{(l)}_{,\delta},\\
{\bf N}^{kl} &=& {\cal F}^{(k)}_{,\alpha} B^{\alpha\beta} {\cal F}^{(l)}_{,\beta}
= \bar{\cal F}^{(k)}_{,\alpha} B^{\alpha\beta} \bar{\cal F}^{(l)}_{,\beta}
,\\
{\bf V}^k &=& {\cal F}^{(k)}_{,\alpha} B^{\alpha\beta} V_{,\beta}
= \bar{\cal F}^{(k)}_{,\alpha} B^{\alpha\beta} V_{,\beta} ,
\end{eqnarray}
and ${\bf W}^{kI}= 0$,
which means that
the spurious components of one-body operators $F^{(k)}$
do not play any role in the projected LHA formalism.
All we need to do is to keep $q^I$ constant all the way
along the collective path.
Then, we can use the original set of one-body operators $\{ F^{(k)} \}$
on which the LHA equations are projected, without modifying
Eqs.~(\ref{projected_LHA}) to (\ref{projected_LHA_V}).
In the classical theory of the P+Q model discussed in the next section
the constants of motion are coordinates $q^I$, and not momenta $p_I$.
Therefore this is exactly the case where the mixing of spurious components
does not play any role.
For further details, see the discussion around Eq.~(\ref{PQ_spurious}).
\section{Collective coordinates for the P+Q model}
\label{sec: P_plus_Q}
In this section,
we investigate the structure of the self-consistent collective coordinates
(cranking operators) at the Hartree-Bogoliubov (HB) state
for the pairing-plus-quadrupole (P+Q) Hamiltonian,
as a first step towards the large amplitude collective motion
in heavy nuclei.
We also try to find a set of one-body operators which can approximate the
RPA eigenvectors.
\subsection{The quasiparticle RPA for the P+Q model}
We apply the symplectic version of LHA with the extended adiabatic
approximation to the P+Q Hamiltonian.
At the HB state, as shown in Sec.~\ref{sec: SLHA},
the force condition can be always satisfied and the local RPA equation
is equivalent to the quasiparticle RPA for the constrained Hamiltonian
\begin{eqnarray}
H' &=& H - \sum_{\tau=n,p} \lambda_\tau N_\tau ,\\
H &=& \sum_k \epsilon_k c_k^\dagger c_k
-\sum_{\tau=n,p} \frac{G_\tau}{2}
\left( P_\tau^\dagger P_\tau + P_\tau P_\tau^\dagger \right)
-\frac{\chi}{2} \sum_{K=-2}^2 Q_{2K}^\dagger Q_{2K} \\
&=& \sum_k \epsilon_k c_k^\dagger c_k
-\frac{1}{2} \sum_\sigma \kappa_\sigma R_\sigma R_\sigma
+\frac{1}{2} \sum_\sigma \kappa_\sigma S_\sigma S_\sigma ,
\end{eqnarray}
where $\epsilon_k$ are spherical single-particle energies and
$N_\tau = \sum_{k\in\tau} c_k^\dagger c_k$ are the number operators
for neutrons ($\tau=n$) and protons ($\tau=p$).
The operators
$R_\sigma$ and $S_\sigma$ are the hermitian and anti-hermitian
components, respectively, of the pairing operators,
$P_\tau^\dagger=\sum_{k\in\tau, k>0} c_k^\dagger c_{\bar k}^\dagger$,
and the dimensionless quadrupole operators,
$Q_{2K} = b_0^{-2} \sum_{kl} \bra{k} r^2 Y_{2K} \ket{l} c_k^\dagger c_l$,
where $b_0=(\hbar/m\omega_0)^{1/2}$ is the harmonic oscillator length.
The Hamiltonian contains five operators of $R_\sigma$-type and four of
$S_\sigma$-type.
Together with the corresponding coupling constants $\kappa_\sigma$,
they are given by
\begin{equation}
\label{PQ_table}
\begin{array}{ccccccc}
R_\sigma &=& (P_+^{(+)})_n, & (P_+^{(+)})_p,
&Q^{(+)}_{20}, &Q^{(-)}_{21}, &Q^{(+)}_{22},\\
S_\sigma &=& (P_-^{(+)})_n, & (P_-^{(+)})_p,& &Q^{(+)}_{21}, &Q^{(-)}_{22},\\
\kappa_\sigma &=& G_n, & G_p, &\chi, &\chi, &\chi,
\end{array}
\end{equation}
where
\begin{eqnarray}
(P_\pm^{(+)})_\tau &=& \frac{1}{\sqrt{2}} ( P_\tau \pm P_\tau^\dagger ),
\quad \mbox{ for } \tau=n,p, \nonumber \\
\label{sig_good_Q}
Q^{(\pm)}_{2K} &=& \frac{1}{\sqrt{2}} ( Q_{2K} \pm Q_{2-K} ) , \quad
\mbox{ for } K=0,1,2.
\end{eqnarray}
The signs $(\pm)$ indicate the signature quantum number,
$e^{-i\pi J_x} O^{(\pm)} e^{i\pi J_x} = \pm O^{(\pm)}$.
Following the standard formulation of the model,
we will neglect the Fock terms,
the contributions of the pairing force to the Hartree potential
and those of the quadrupole force to the pairing potential
\cite{Bel59,KS60,Bes61,KS63,BK_I,BK_II,BK_III,BK_IV,BK_V}.
After minimising the HB total energy
and diagonalising the HB matrix,
we obtain
the HB ground state $\ket{\Psi_0}$ and the quasiparticle energies $E_k$.
Using $\ket{\Psi_0}$ as the reference state,
the classical Hamiltonian is calculated as
\begin{eqnarray}
\label{H'_PQ}
{\cal H}'(\xi,\pi) &=& {\cal H}(\xi,\pi)
- \sum_\tau \lambda_\tau {\cal N}_\tau (\xi,\pi) \nonumber\\
&=& E_0' + \sum_k E_k \bar\rho_{kk}
- \frac{1}{2} \sum_\sigma \kappa_\sigma
\left( \sum_{i>j} R_\sigma(ij) (\bar\kappa_{ij}^* + \bar\kappa_{ij})
+\sum_{ij} R'_\sigma(ij) \bar\rho_{ji} \right)^2 \nonumber \\
&&\hspace*{2.5cm} +\frac{1}{2} \sum_\sigma \kappa_\sigma
\left( \sum_{i>j} S_\sigma(ij) (\bar\kappa_{ij}^* - \bar\kappa_{ij})
+\sum_{ij} S'_\sigma(ij) \bar\rho_{ji} \right)^2 .
\end{eqnarray}
Applying the Holstein-Primakoff mapping (\ref{H_P_mapping})
and taking into account that the $R'(ij)$ are symmetric and
$S'(ij)$ are anti-symmetric with respect to the indices $(ij)$,
one can see that $\bra{\Psi(t)} R_\sigma \ket{\Psi(t)}$
($\bra{\Psi(t)} S_\sigma \ket{\Psi(t)}$) are real (pure imaginary),
and indeed the Hamiltonian (\ref{H'_PQ}) is real.
Differentiating ${\cal H}'$ with respect to $\pi$ and $\xi$,
the mass and curvature parameters at the HB state, defined by
Eqs.~(\ref{EAT_HFB}) and (\ref{SLHA_HFB}), are given by
\begin{eqnarray}
\label{B_PQ}
\widetilde{B}^{\alpha\beta}(\xi=\pi=0) &=&
B^{\alpha\beta} - \sum_{\tau=n,p} \lambda_\tau f^{(1)N_\tau \alpha\beta}
\nonumber \\
&=& E_\alpha \delta_{\alpha\beta}
- 2\sum_\sigma \kappa_\sigma S_\sigma(\alpha) S_\sigma(\beta) ,\\
\label{V_PQ}
\widetilde{V}_{;\alpha\beta}(\xi=\pi=0) &=&
V_{,\alpha\beta} - \sum_{\tau=n,p} \lambda_\tau f^{N_\tau}_{,\alpha\beta}
\nonumber \\
&=& E_\alpha \delta_{\alpha\beta}
- 2\sum_\sigma \kappa_\sigma R_\sigma(\alpha) R_\sigma(\beta) .
\end{eqnarray}
Here, again the Greek index indicates a pair of 2qp index,
so that $R(\alpha)$ are the 2qp matrix elements $R(\alpha)= R(ij)$ and
$E_\alpha$ are the 2qp energies $E_\alpha = E_i+E_j$.
We see that the terms $R'(ij)\bar{\rho}_{ji}$ and $S'(ij)\bar{\rho}_{ji}$
in the Hamiltonian do not contribute to
the mass and curvature parameters, which is consistent with the fact
that the terms proportional to $(a^\dagger a)$ in one-body operators
are neglected in the quasiparticle RPA.
Solving the RPA is equivalent
to diagonalisation of the matrix
$\widetilde{B}^{\alpha\gamma} \widetilde{V}_{;\gamma\beta}$.
For separable forces such as the P+Q model,
the problem of diagonalisation can be reduced to a root search
in a multi-dimensional dispersion equation,
which facilitates the numerical calculations for heavy nuclei \cite{RS80}.
If the ground state has an axial symmetry, the RPA matrix is block-diagonal
and can be divided according to $K$ quantum numbers (eigenvalues of $J_z$).
However, as we will show below, some nuclei have triaxial HB ground states
for which $K$ is no longer a good quantum number.
Thus, we use the signature quantum number $r$,
which is good even for triaxial cases, and
the RPA dispersion equation can be divided
into positive ($r=+1$) and negative ($r=-1$) signature sectors.
The construction of the response functions
and the coupled dispersion equations
can be found, for example, in Ref. \cite{NMMS96}.
\subsection{Numerical results}
\label{sec: results}
We follow the second and third of the series of papers by Baranger and Kumar
\cite{BK_II,BK_III} for details of the P+Q model.
The model space consists of two major shells of $N_{\rm osc} = 5, 6$
and $4, 5$ for neutrons and protons, respectively.
Different harmonic oscillator frequencies are used for neutrons
and protons in order to make the root mean square radii the same \cite{BK_II}.
The single-particle energies
are taken from Table~1 in Ref. \cite{BK_III},
where they are listed in units of the
harmonic oscillator frequency $\hbar\omega_0=41.2A^{-1/3}$ MeV.
The pairing force strengths are
$G_n=22A^{-1}$ MeV and $G_p=27A^{-1}$ MeV
for neutrons and protons, respectively.
The quadrupole force strength is $\chi=70 A^{-1.4}$ MeV.
The effective charge in the $E2$ operator is taken as $e_n = 1.5 Z A^{-1}$
and $e_p = 1 + e_n$ for neutrons and protons, respectively.
Since matrix elements of the quadrupole force in upper shells are known
to be too strong \cite{BK_II},
we modify the quadrupole operators
by multiplying all quadrupole matrix
elements by a factor
$\zeta = (N_{\rm L}+\frac{3}{2})/(N_{\rm osc}+\frac{3}{2})$,
where $N_{\rm osc}$ is the oscillator quantum number of the shell under
consideration and $N_{\rm L}$ is that of the lower major shell.
Furthermore,
due to the difference of harmonic oscillator frequencies
for neutrons and protons,
the operators are multiplied by factors $\alpha_n^2=(2N/A)^{2/3}$
and $\alpha_p^2=(2Z/A)^{2/3}$ for neutrons and protons, respectively.
Thus,
the dimensionless quadrupole operators in the interaction (\ref{PQ_table})
are actually defined by
\begin{eqnarray}
\label{modify_Q}
Q_{2K} &=& (Q_{2K})_n + (Q_{2K})_p ,\nonumber\\
(Q_{2K})_n &=& \alpha_n^2 \zeta b_0^{-2} (r^2 Y_{2K})_n , \quad
(Q_{2K})_p = \alpha_p^2 \zeta b_0^{-2} (r^2 Y_{2K})_p .
\end{eqnarray}
These operators will be referred to simply as the ``quadrupole operators''
in the following discussion.
We have performed calculations for even-even nuclei
in the rare-earth region ($N=82\sim 126$, $Z=50\sim 82$)
and have investigated the structure of RPA eigenvectors.
Since $\beta$ and $\gamma$ vibrations in these nuclei were studied by
Baranger and Kumar in Ref. \cite{BK_III},
we can compare our microscopic RPA results
with their semi-microscopic ones.
In Table~\ref{HFB_plus_RPA}, results of the HB and the RPA
calculations are presented.
The equilibrium deformations ($\beta$, $\gamma$, $\Delta_n$, $\Delta_p$)
at HB ground states are found to agree with Table~2 in Ref. \cite{BK_III},
except $^{190}$Pt which has a small triaxiality ($\gamma=52.6^\circ$)
in our result but has a oblate shape ($\gamma=60^\circ$) in theirs.
The deformation parameters, $\beta$ and $\gamma$, of the HB states
are defined by \cite{BK_II,BK_III},
\begin{equation}
\chi \bra{\Psi_0} Q_{20}^{(+)} \ket{\Psi_0}
= \hbar\omega_0 \beta \cos\gamma ,\quad
\chi \bra{\Psi_0} Q_{22}^{(+)} \ket{\Psi_0}
= \hbar\omega_0 \beta \sin\gamma .
\end{equation}
We have solved the RPA in the positive signature sector where
the spurious modes, the pairing and spatial rotation, are coordinates
(not momenta) and are characterised
by zero eigenmodes of the mass matrix,
$B^{\alpha\beta} f^I_{,\beta} = 0$.
For spherical nuclei ($\beta=0$), the lowest excited states
with multipolarity $\lambda=2$ are selected
as quadrupole vibrations and presented in columns 7 and 8.
For prolate ($\beta>0,\gamma=0$) and oblate ($\beta<0,\gamma=0$) deformed
cases, the first excited states (except zero spurious modes)
with $K=0$ and $K=2$ are
taken as $\beta$ and $\gamma$ vibrations, respectively.
However, for some cases, the lowest excited states appear to be
non-collective and we have selected higher-energy states.
These states are indicated with $*$ in the table.
For a triaxial ground state, the first excited states are selected and
the excitation energies are shown in column 7 while the transition
amplitudes are given in columns 8 and 10.
The $E2$ amplitudes $M(E2)_K$ are intrinsic values and calculated as
\begin{equation}
M(E2)_K =
\left| \bra{n^{(+)}} e_n \zeta (r^2Y_{2K})^{(+)}_n
+ e_p \zeta (r^2Y_{2K})^{(+)}_p \ket{0} \right| ,
\end{equation}
where $\ket{0}$ and $\ket{n^{(+)}}$ are the RPA ground and excited states
with signature $r=+1$, respectively.
For spherical nuclei, states with $K=0$, 1 and 2
are degenerate in energy and we select the $K=0$ one for $\ket{n^{(+)}}$.
For triaxial shapes, the state $\ket{n^{(+)}}$
has non-zero $E2$ amplitudes for both $K=0$ and $K=2$ which are shown
in columns 8 and 10 respectively.
First we compare the RPA frequencies with the
harmonic formula given by Baranger and Kumar,
Eqs.~(13), (14) and (15) in Ref. \cite{BK_III}.
Unlike our results,
this formula contains a core contribution to the collective mass
(which is denoted by $B^{\rm c}$).
Since our RPA calculations are carried out in the same model space,
we should compare the RPA frequencies with those calculated
by setting $B^{\rm c}=0$.
These are shown in Fig.~\ref{RPA_BK_omega}.
Open symbols are the results of the harmonic formula while solid ones are
the RPA frequencies.
For spherical and triaxial nuclei, the same RPA frequencies are displayed
in the figure for $\beta$ and $\gamma$ vibrations.
The figure shows that the frequencies obtained by the formula are
larger than the RPA frequencies.
For $\gamma$ vibrations, the isotope dependence is well reproduced,
even though the absolute values are mostly overestimated by up to a factor two.
For $\beta$ vibrations,
the harmonic formula breaks down,
especially in well-deformed nuclei.
It fails to reproduce both the isotope dependence and absolute values.
Let us try to understand the origin of the harmonic formula.
It can be derived by taking the small-amplitude limit of
the adiabatic time-dependent theory of collective motion
developed by Baranger and Kumar \cite{BK_IV,BK_V}
which assumes that the collective coordinates are the expectation values
of the mass quadrupole operators
(more precisely the deformation parameters of the Nilsson potential).
In contrast with the formula,
the RPA normal-mode coordinates are self-consistently determined by
diagonalising the RPA matrix.
Therefore, the difference between open and closed symbols seen
in the figure can be attributed primarily to the difference
in the collective coordinates.
Although it is not the purpose of this paper to reproduce the
experimental data, we also show observed excitation energies of
$\beta$ and $\gamma$ bands for axially deformed nuclei ($Z=60\sim 74$)
in Fig.~\ref{RPA_BK_omega} \cite{SHS91}. Agreement with the experimental
data are significantly improved by choosing the collective coordinates
properly. One can see that the isotope dependence is roughly reproduced
in this model. However, we still overestimate the energies by typically
about 500 keV, which is at least partly due to the neglect of
anharmonic effects, and may point at the limitations of the simple model.
In Fig.~\ref{RPA_E2}, we show $E2$ transition amplitudes for the RPA
states.
Evidently the $E2$ amplitudes are smaller for $\beta$ vibrations than
$\gamma$ vibrations.
This is partly because the $\beta$ vibrations are less collective, but
also because the collective coordinates for $\beta$ vibrations are
more complex than those for $\gamma$ motion (see below).
Now let us analyse the structure of the RPA normal-modes by
using the projection techniques described in Sec.~\ref{sec: PLHA},
using various sets of one-body operators.
First we utilise elementary multipole operators $\{ F^{(k)} \}$
on which the RPA matrix is projected.
We adopt monopole operators (rank-0),
quadrupole operators with radial dependence $r^0$ and $r^2$
(rank-2, $K=0,2$),
those with spin dependence (rank-2, $K=0,2$), and
hexadecapole operators (rank-4, $K=0,2$):
\begin{eqnarray}
\label{PRPA_op_1}
F^{(k)}=&& (P_+)_\tau, (P_-)_\tau,
(Q_{20})_\tau, (Q_{22})_\tau,
(r^0Y_{20})_\tau, (r^0Y_{22})_\tau, \nonumber\\
&& ([r^0Y_2 \times {\bf s} ]^{(2)}_{K=0})_\tau,
([r^0Y_2 \times {\bf s} ]^{(2)}_{K=2})_\tau,
([r^2Y_2 \times {\bf s} ]^{(2)}_{K=0})_\tau,
([r^2Y_2 \times {\bf s} ]^{(2)}_{K=2})_\tau, \nonumber \\
&& (r^4 Y_{40})_\tau, (r^4 Y_{42})_\tau , \quad \tau = n,p,
\end{eqnarray}
where all operators have the positive signature $r=+1$,
as in Eq.~(\ref{sig_good_Q}), but the superscript
$(+)$ denoting the signature quantum number is suppressed from here on.
All matrix elements of rank-2 operators are multiplied by
factors $\zeta$ and $\alpha_\tau^2$,
just as for the quadrupole operators (\ref{modify_Q}).
For $\beta$ vibrations, there are fourteen relevant operators of $K=0$
while there are ten operators of $K=2$ for $\gamma$.
For a spherical case, the number reduces to eight (only rank-2 operators).
For a triaxial case, all 24 operators can mix together.
The calculated frequencies $\hbar\bar{\Omega}$ of the projected RPA equation
(\ref{projected_LHA})
are shown in Fig.~\ref{PRPA_omega} as
open symbols with dashed lines and are
compared with the real RPA frequencies (solid symbols).
As is clearly seen, the projection onto the operators (\ref{PRPA_op_1})
fails to reproduce
the excitation energies both for $\beta$ and $\gamma$ vibrations.
Not only the absolute values but also the isotope dependence turns out to be
incorrect for all nuclei (except for $^{208}$Pb).
The result indicates
that it is very difficult to obtain sensible approximation
by using elementary one-body operators.
This is mainly due to the fact that the projected RPA eigenvectors have
unrealistically large amplitudes for high-lying 2qp components.
In Ref. \cite{NWD99},
we have demonstrated for a few Sm isotopes that this can be remedied
by introducing a cut-off energy
for the 2qp matrix elements.
For $^{208}$Pb, which is spherical and has $\Delta_n=\Delta_p=0$,
the projection accidentally works well because
the set of available 1p1h modes is very limited,
$\nu(i_{11/2}(i_{13/2})^{-1})$, $\nu(g_{9/2}(i_{13/2})^{-1})$,
$\pi(h_{9/2}(h_{11/2})^{-1})$, $\pi(f_{7/2}(h_{11/2})^{-1})$,
all of which have almost the same particle-hole excitation energies.
In order to check the effect of suppressing the high-energy components,
we also perform the projected RPA calculation
with the state-dependent operators whose 2qp matrix elements are
weighted with a factor $(E_{\rm 2qp})^{-2}$.
This means that we employ a set of hermitian
one-body operators $\{ \widetilde{F}^{(k)}\}$ defined by
\begin{equation}
\widetilde{F} \equiv \sum_\alpha
\frac{F(\alpha)}{(E_\alpha)^2} (a^\dagger a^\dagger)_\alpha
+ \mbox{h.c.},
\label{scale}
\end{equation}
where $\alpha$ indicates the 2qp index and
$F(\alpha)=\bra{\alpha} F \ket{0}$.
This suppression factor $(E_{\rm 2qp})^{-2}$ can be derived analytically
as follows
if the residual Hamiltonian consists only of a single-mode separable force
$H=-\frac{1}{2} \kappa R R$ and the RPA frequency is much smaller
than 2qp energies.
In the single-mode case,
from Eqs.~(\ref{B_PQ}) and (\ref{V_PQ}),
the RPA equation becomes
\begin{equation}
\left( (E_\alpha)^2 - (\hbar\Omega)^2 \right) f_{,\alpha}
= 2\kappa R(\alpha) \sum_\beta E_\beta R(\beta) f_{,\beta} .
\end{equation}
Thus, we can analytically determine the RPA
eigenvectors $f_\alpha$ as
\begin{equation}
f_{,\alpha} \propto \frac{R(\alpha)}{(E_\alpha)^2 - (\hbar\Omega)^2} .
\end{equation}
In the limit of $\hbar\Omega \ll E_\alpha$, this leads to the quoted
suppression factor $(E_{\rm 2qp})^{-2}$.
In the calculation, we adopt only the operators in the separable
interactions:
\begin{equation}
\label{PRPA_op_2}
\widetilde{F}^{(k)}=
(\widetilde{P}_+)_\tau, (\widetilde{P}_-)_\tau,
(\widetilde{Q}_{20})_\tau, (\widetilde{Q}_{22})_\tau,
\quad \tau = n,p.
\end{equation}
This results in a six dimensional projected RPA matrix for $\beta$
vibrations and in two dimensional one for $\gamma$ vibrations and
for spherical cases.
Even for triaxial cases, the matrix dimension is thus only eight.
The results are shown in Fig.~\ref{PRPA_omega} by open symbols with
solid lines.
The results are dramatically improved from the previous ones and
the obtained frequencies are almost comparable to those of RPA
for both $\beta$ and $\gamma$ modes of all nuclei.
We can also see the significant improvement over results of
the harmonic formula, especially for the $\beta$
vibrations (Fig.~\ref{RPA_BK_omega}).
This is because we incorporate the monopole pairing operators in
Eq.~(\ref{PRPA_op_2}) in addition to the quadrupole ones.
It is worth noting that the frequency of the projected RPA gives an
upper limit to the real RPA frequency because it is obtained by solving
the RPA in a restricted space.
On the other hand, this is not necessarily true for the harmonic formula
of Ref.~\cite{BK_III}.
Let us examine the eigenvectors of projected RPA equation in more detail.
For spherical nuclei, the RPA eigenvector has a good isoscalar character
and can be well approximated by
\begin{equation}
\bar{f}^Q = c_n (\widetilde{Q}_2)_n + c_p (\widetilde{Q}_2)_p ,
\end{equation}
with $c_n \approx c_p$.
Here the tilde indicates that the matrix elements
include a suppression factor as in Eq.~(\ref{scale}).
A violation of this isoscalar character is seen for
$^{138}$Ba, $^{140}$Ce, $^{142}$Nd and Pb nuclei
in which either the neutrons or protons pairing gap vanishes.
The component with finite pairing gap is found to be larger
than the one with zero gap,
so that $c_n > c_p$ for Pb and $c_n < c_p$ for
$^{138}$Ba, $^{140}$Ce and $^{142}$Nd.
For deformed nuclei, where the collectivity of the
vibrational states is smaller than for spherical nuclei and the pairing
modes can mix with the quadrupole ones, the situation is more complex.
In Tables~\ref{PRPA_Dy} and \ref{PRPA_Os},
taking Dy and Os isotopes as examples,
we show the coefficients of the approximate eigenvectors,
$\bar{f} = \sum_k c_k \widetilde{F}^{(k)}$,
with respect to the set of operators in Eq.~(\ref{PRPA_op_2}).
The $\gamma$ vibrations approximately have isoscalar character
and the collective coordinate can be approximated by
\begin{equation}
\bar{f}^\gamma = c_n (\widetilde{Q}_{22})_n + c_p (\widetilde{Q}_{22})_p ,
\end{equation}
with $c_n \approx c_p$.
For the $\beta$ vibration, no such simple form can be found, since
the properties of RPA eigenmodes change from one nucleus to the other.
The $\beta$ vibrations for $^{156-162}$Dy possess a quadrupole nature,
$c_n (\widetilde{Q}_{20})_n + c_p (\widetilde{Q}_{20})_p$,
with $c_n > c_p$,
although there are substantial contributions from the monopole pairing modes.
For $^{164}$Dy, there is a dominant pairing component $(P_+)_n$.
For Os isotopes, $^{182\sim 186}$Os have dominant proton pairing $(P_+)_p$,
while $^{192}$Os has a quadrupole nature ($c_n < c_p$).
For the triaxial nuclei $^{188,190}$Os,
although the $K=0$ and $K=2$ components are mixed,
the vibrational state is of
isoscalar quadrupole nature with $c_n \approx c_p$.
The changes observed in the RPA normal-mode coordinates
for the $\beta$
vibrations are difficult to be systematically summarised and
turn out to be very sensitive to details of the underlying shell structure.
In order to confirm the importance of the monopole components
for $\beta$ vibrations,
we have performed a calculation using only the quadrupole operators
weighted with the factor $E_{\rm 2qp}^{-2}$,
$(\widetilde{Q}_{20})_\tau$ and $(\widetilde{Q}_{22})_\tau$
($\tau=n,p$).
For spherical nuclei, there is no change from the previous results
since the monopole and quadrupole components are exactly decoupled.
For deformed nuclei only, the results are shown in Fig.~\ref{PRPA_omega}
by open symbols with dotted lines.
Differences between solid and dotted lines come from the contributions
of monopole pairing operators.
The figure clearly shows that the normal-mode coordinate for $\beta$
vibrations has a significant amount of monopole components.
We now wish to quantify the quality of the approximation.
One measure is the difference between
the real RPA frequencies $\hbar\Omega$ and
the projected RPA $\hbar\bar{\Omega}$,
but as usual measures based on the eigenvectors are more sensitive.
We define two criteria to directly check the closeness between
the eigenvectors $f$ and $\bar{f}$.
Choosing the normalisation (of $f$ and $\bar{f}$)
$f_{,\alpha} B^{\alpha\beta} f_{,\beta} =
\bar{f}_{,\alpha} B^{\alpha\beta} \bar{f}_{,\beta} = 1$,
those are calculated as
\begin{eqnarray}
\label{delta_B}
\delta_B &=&
f_{,\alpha} B^{\alpha\beta} ( f_{,\beta}-\bar{f}_{,\beta} )
=1 - f_{,\alpha} B^{\alpha\beta} \bar{f}_{,\beta} ,\\
\delta_1 &=& 1-\frac{(f,\bar{f})}
{\sqrt{(f, f)(\bar{f},\bar{f})}} ,
\end{eqnarray}
where $(f,f')=\sum_\alpha f_{,\alpha} f'_{,\alpha}$.
We have $0 \leq \delta \leq 1$ with
$\delta=0$ corresponding to the exact projection and $\delta=1$ to the
case where $\bar{f}$ is orthogonal to $f$.
The values of $\delta_B$ and $\delta_1$ are listed in
Tables~\ref{PRPA_Dy} and \ref{PRPA_Os}.
From the values of $\delta_B$,
we see that the projection works reasonably well
for Dy and Os isotopes.
The worst case is the $\beta$ vibration of $^{192}$Os for which
$\delta_B\approx 0.3$.
Roughly speaking, the one-body operators (\ref{PRPA_op_2}) possess
70\% of overlap with the real eigenvectors for $^{192}$Os
($\delta_B\approx 0.3$),
and more than 90\% for the others ($\delta_B \lesssim 0.1$).
The values of $\delta_1$ for $\beta$ vibrations are often much larger than
the corresponding $\delta_B$
(see $^{156,158}$Dy and triaxial nuclei $^{188,190}$Os).
This indicates that the $\bar{f}$ for $\beta$ contains some spurious
components.
In the present calculations,
the pairing and spatial rotations are the obvious spurious motions
which are zero eigenmodes of the mass matrix,
\begin{equation}
\label{PQ_spurious}
B^{\alpha\beta} f^I_{,\beta} = 0 , \quad\mbox{for}\quad
f^I = N_n, N_p, J_x .
\end{equation}
The measure $\delta_B$ (\ref{delta_B}) is insensitive to admixture of these
spurious components in $\bar{f}$.
On the other hand, the value of $\delta_1$ will increase by mixing in
spurious components.
As we have discussed in the last part of Sec.~\ref{sec: PLHA},
the spurious modes $f^I$ do not affect the projected LHA equations.
Therefore, the large values of $\delta_1$ should not cause serious problems
in application of the projected LHA to the large amplitude collective motion.
In Fig.~\ref{PRPA_delta}, the values of $\delta_B$ are shown for
the projected RPA using the set of elementary operators (\ref{PRPA_op_1})
(represented by dashed lines) and the set of state-dependent operators
containing the factor $E_{\rm 2qp}^{-2}$ (\ref{PRPA_op_2}) (solid lines).
For $\gamma$ vibrations, the suppression factor $E_{\rm 2qp}^{-2}$
can improve the quality of projection by an order of magnitude
and we have $\delta_B \lesssim 0.1$ for all nuclei.
For $\beta$ vibrations, there are some nuclei in the range $Z=68\sim 80$
for which the operators in Eq.~(\ref{PRPA_op_2}) are not enough to produce
the good approximation.
This confirms the complexity of the normal-mode coordinates for $\beta$.
We also see a strong isotope dependence in Fig.~\ref{PRPA_delta}a.
For instance, the Er isotopes show a staggering behaviour in $\delta_B$.
This behaviour is found to be
closely related to the collectivity of the states.
Since the $\beta$ vibrations contains the pairing collectivity in addition to
the quadrupole one,
the $E2$ amplitudes are not necessarily the best indicator of the
collectivity of the states.
We show in Fig.~\ref{PRPA_stag} a following quantity for the RPA eigenmodes
of $\beta$ vibrations in Er nuclei \cite{AKL81}:
\begin{equation}
\label{Neff}
N_{\rm eff}^\mu \equiv
\frac{\sum_\alpha ( \hbar\Omega f^\mu_{,\alpha}f^\mu_{,\alpha}
+ (\hbar\Omega)^{-1} g^\alpha_{,\mu} g^\alpha_{,\mu} )}
{\sum_\beta f^\mu_{,\beta} g^\beta_{,\mu}
(\hbar\Omega f^\mu_{,\beta}f^\mu_{,\beta}
+(\hbar\Omega)^{-1} g^\beta_{,\mu} g^\beta_{,\mu})} ,
\quad \mbox{(no summation with respect to $\mu$).}
\end{equation}
If $n$ 2qp components equally contribute to the RPA mode, we have
$N_{\rm eff}=n$.
Thus, Eq. (\ref{Neff}) gives an effective number of 2qp excitations
of which the RPA eigenmode consists.
The figure shows that the collectivity of $\beta$ vibrations in Er isotopes
is weak (compared to the $\gamma$ vibrations for which
$N_{\rm eff} \gtrsim 8$ in most cases),
and $N_{\rm eff}$ shows a staggering pattern correlated with that for
$\delta_B$ (while the $E2$ amplitudes in Fig.~\ref{RPA_E2} show no sign of
staggering behaviour).
For $^{164,168}$Er, there are dominant 2qp components in the RPA eigenmodes,
which decrease $N_{\rm eff}$ and increase $\delta_B$.
For those cases where $\delta_B \gtrsim 0.3$ even with state-dependent
operators, the $\beta$ vibrations are not collective at all,
$N_{\rm eff} \lesssim 3$.
\section{conclusion and discussion}
\label{sec: conclusions}
We have examined different version of LHA techniques
to see whether the conservation laws in the quantum level are
satisfied in the classical LHA.
It turns out that the restriction to point transformations must
be lifted in order that spurious motion leads to the exact zero modes.
The symplectic version of the LHA, which includes an extended adiabatic
approximation,
can guarantee the exact separation of zero-energy spurious modes.
The projected LHA can be used to truncate the local RPA calculations.
We have shown that there is no problem with spurious modes
for the projection method, at least
when the spurious modes are zero eigenmodes of the mass matrix.
This is exactly the case for the pairing and spatial rotation in the
P+Q model for which
we have investigated the structure of the collective coordinates
for quadrupole vibrations
and examined the possibility of expressing the self-consistent
cranking operator in terms of a limited number of one-body operators.
It seems very difficult to approximate the normal-mode vectors
with only elementary one-body operators.
The difficulty disappears, however,
when we use a small number of {\it state-dependent} one-body operators.
At HB equilibrium, the collective coordinates for quadrupole vibrations
in spherical nuclei and for $\gamma$ vibrations ($K=2$) in axially deformed
nuclei can be roughly approximated as $f\approx \widetilde{Q}_{2K}$, where
$\widetilde{Q}_{2K}$ is a state-dependent ``mass'' quadrupole operator
defined by Eq.~(\ref{modify_Q}) weighted by a factor
$E_{\rm 2qp}^{-2}$ for each 2qp matrix element (\ref{scale}).
However, for $\beta$ vibrations in axial deformed nuclei and
for all excitations in triaxial nuclei,
the monopole pairing components are as important as the quadrupole ones.
It is worth emphasising that the collective coordinates are
very different from the usual (state-independent) mass quadrupole
operator even for $\gamma$ and spherical cases.
This shows the importance of a self-consistent determination of
the collective coordinates for large amplitude collective motion,
because the coordinates are now found to have a strong state-dependence as well.
The structure of the self-consistent cranking operators is clearly
changing when we move from spherical to axially deformed and triaxial
nuclei.
For the study of large amplitude
collective motion in heavy nuclei for which
the diagonalisation of the RPA matrix becomes too time-consuming,
the results of this paper gives a suggestion for a
choice of a state-dependent basis of operators.
The choice of a limited set of (state-dependent) basis operators
provides a practical way to solve the LHA through the projection.
With the use of self-consistent cranking operators,
the LHA should provide a significant improvement over a conventional
CHFB calculation based on fixed cranking operators.
Despite all the above qualities,
the symplectic version of LHA has a difficulty, namely
the determination of $f^\mu_{,\alpha\beta}$,
the second derivatives of the final
coordinates $q^\mu$ with respect to the original coordinates $\xi^\alpha$.
For the extended adiabatic transformation (\ref{EAT_0}),
we also need to calculate $f^{(1)\mu\alpha\beta}$,
the terms of second order in $\pi$.
At HB states, we need these quantities for spurious coordinates $\mu=I$ only,
$f^I_{,\alpha\beta}$ and $f^{(1)I\alpha\beta}$,
which can be calculated
because we know {\it a priori} the one-body operators corresponding to
the spurious motions.
We have done this for P+Q model.
Away from equilibrium, however, in addition to these $\mu=I$,
we need the second derivatives of the collective coordinate, $\mu=1$.
This is not trivial because
the local RPA equation determines only the first derivatives,
$f^1_{,\alpha}$ and $g^\alpha_{,1}$.
Unfortunately, so far we do not have a general method to
determine $f^1_{,\alpha\beta}$ and $f^{(1)1\alpha\beta}$.
The projection method may actually provide us with a partial answer
\cite{WKD90}.
If we project the LHA equation onto a set of one-body operators
$\{ F^{(k)} \}$,
the equation provides us with a local representation of
collective coordinate in terms of a linear combination of the $F^{(k)}$'s.
Using the explicit form
\begin{equation}
F^{(k)} = \sum_{i > j} F^{(k)}(ij)
(a_i^\dagger a_j^\dagger + a_j a_i )
+ \sum_{ij} F^{'(k)}(ij) a_i^\dagger a_j ,
\end{equation}
we find
\begin{equation}
\label{f_alpha}
f^1_{,\alpha} \equiv \frac{\partial q^1}{\partial \xi^\alpha}
= \sum_k c_k {\cal F}^{(k)}_{,\alpha}
= \sqrt{2} \sum_k c_k F^{(k)}(\alpha) ,
\end{equation}
where $F^{(k)}(\alpha)=F^{(k)}(ij)$ for 2qp index $\alpha=(ij)$ and
we assume that the operators $F^{(k)}$ are hermitian.
Here we use a local coordinate system, in which the point
under consideration is the origin of the phase space,
and the relation (\ref{2qp_R}) to obtain Eq.~(\ref{f_alpha}).
Since we know the explicit form of the operators $F^{(k)}$,
we are now able to calculate the second derivatives $f^1_{,\alpha\beta}$ and
second-order terms in momentum $f^{(1)1\alpha\beta}$.
\begin{eqnarray}
\label{f_ab}
f^1_{,\alpha\beta} &\equiv&
\frac{\partial^2 q^1}{\partial\xi^\alpha\partial\xi^\beta}
\approx \sum_k c_k {\cal F}^{(k)}_{,\alpha\beta}
= 4 \sum_k c_k G^{(k)}(\alpha\beta) ,\\
\label{f_1ab}
f^{(1)1\alpha\beta} &\equiv&
\frac{\partial^2 q^1}{\partial\pi_\alpha\partial\pi_\beta}
\approx \sum_k c_k \frac{\partial^2{\cal F}^{(k)}}
{\partial\pi_\alpha\partial\pi_\beta}
= 4 \sum_k c_k G^{(k)}(\alpha\beta) ,
\end{eqnarray}
where we neglect the derivative of $c_k$, for simplicity.
$G^{(k)}(\alpha\beta)$ are related to $F^{'(k)}(ij)$, by
\begin{equation}
G^{(k)}(\alpha\beta) \equiv F^{'(k)}(ij) \delta_{kl} , \quad
\mbox{for }\alpha=(ik), \beta=(jl).
\end{equation}
Although, following this procedure, it is possible to calculate
$f^1_{,\alpha\beta}$ and $f^{(1)1\alpha\beta}$,
we have to note that, unlike the determination of $f^1_{,\alpha}$ by means
of the RPA,
the determination of these quantities through Eqs.~(\ref{f_ab}) and
(\ref{f_1ab}) does not depend on the dynamics of the system.
In this approach, part of the dynamics is hidden in the coordinate dependence of the
coefficients $c_k$, which we have ignored for this simple discussion.
In this paper,
we have mainly discussed the structure of the self-consistent
cranking operators for the P+Q model only at HB minimum points.
The structure at points away from the minimum
and the further application of the LHA technique
will be the subject of a future publication.
\acknowledgments
This work was supported by a research grant (GR/L22331) from
the Engineering and Physical Sciences Research Council (EPSRC)
of Great Britain, and through a grant (PN 98.044) from
ALLIANCE, the Franco-British Joint Research programme.
The Laboratoire de Physique Th\'eorique is a Unit\'e Mixte de Recherche
du C.N.R.S., UMR 8627.
We thank Y.~R.~Shimizu for providing us with a numerical code of the RPA
calculation for the pairing-plus-quadrupole forces.
|
\section{Introduction}
The main target of present and future $B$ factories is the study
of CP violation. It will provide tests of the flavor sector of the
Standard Model, which predicts that all CP violation results
from a single complex phase in the quark mixing matrix. The
determination of the sides and angles of the unitarity triangle
plays a central role in this program.\cite{BaBar} The angle
$\beta=\mbox{arg}(V_{td}^*)$ is accessible in a theoretically
clean way through the observation of the CP asymmetry in the decay
$B\to J/\psi K_S$, and a first measurement yielding $\sin 2\beta
=0.79_{-0.44}^{+0.41}$ has just been reported by the CDF
Collaboration.\cite{CDF}
Recently, there has been significant progress in the theoretical
understanding of the hadronic decays $B\to\pi K$, and methods have
been developed to extract information on $\gamma=\mbox{arg}(V_{ub}^*)$
from measurements of the rates for these processes. Here, we discuss
the charged modes $B^\pm\to\pi K$, which are particularly clean from
a theoretical point of view.\cite{us}$^-$\cite{me} For applications
involving the neutral decay modes the reader is referred to the
literature.\cite{FM,Robert}
\section{\boldmath Theory of $B^\pm\to\pi K$ decays\unboldmath}
The hadronic decays $B\to\pi K$ are mediated by a low-energy
effective weak Hamiltonian, whose operators allow for three
distinct classes of flavor topologies: QCD penguins, trees, and
electroweak (EW) penguins. In the Standard Model, the weak couplings
associated with these topologies are known. Data show that the QCD
penguins dominate the decay amplitudes, whereas trees and EW
penguins are subleading and of a similar strength. The theoretical
description of the two charged modes $B^\pm\to\pi^\pm K^0$ and
$B^\pm\to\pi^0 K^\pm$ exploits the fact that the amplitudes for
these processes differ in a pure isospin amplitude, $A_{3/2}$,
given by the matrix element of the isovector part of the effective
Hamiltonian between $B$ and the $(\pi K)$ isospin eigenstate with
$I=\frac 32$. Up to an overall strong-interaction phase $\phi$,
the parameters of this amplitude are determined in the limit of
SU(3) flavor symmetry.\cite{us} SU(3)-breaking corrections can be
calculated in the factorization approximation,\cite{Stech} so that
theoretical uncertainties enter only at the level of nonfactorizable
SU(3)-breaking corrections to a subleading decay amplitude.
A convenient parametrization of the decay amplitudes is
\begin{eqnarray}
{\cal A}(B^+\to\pi^+ K^0) &=& P \Big[ 1
- \varepsilon_a\,e^{i\gamma} e^{i\eta} \Big] \,,\nonumber\\
- \sqrt2\,{\cal A}(B^+\to\pi^0 K^+) &=& P \Big[ 1
- \varepsilon_a\,e^{i\gamma} e^{i\eta}
- \varepsilon_{3/2}\,e^{i\phi}
(e^{i\gamma} - \delta_{\rm EW}) \Big] \,,
\end{eqnarray}
where $P$ is the dominant penguin amplitude defined as the sum of
all terms in the $B^+\to\pi^+ K^0$ amplitude not proportional to
$e^{i\gamma}$, $\eta$ and $\phi$ are strong phases, and
$\varepsilon_a$, $\varepsilon_{3/2}$ and $\delta_{\rm EW}$ are
real hadronic parameters. From a naive quark-diagram analysis,
one does not expect the $B^+\to\pi^+ K^0$ amplitude to receive a
contribution from $b\to u$ tree topologies; however, such a
contribution can be induced through final-state rescattering or
annihilation contributions.\cite{Blok}$^-$\cite{At97} They are
parametrized by $\varepsilon_a$. Model estimates indicate that
$\varepsilon_a$ could be about 5--10\%. In the future, it will be
possible to put upper and lower bounds on this parameter by comparing
the CP-averaged branching ratios for the decays $B^\pm\to\pi^\pm K^0$
and $B^\pm\to K^\pm\bar K^0$.\cite{Fa97} Below, we assume
$\varepsilon_a\le 0.1$; however, our results will be almost
insensitive to this assumption.
The parameter $\delta_{\rm EW}=0.64\pm 0.15$ describes the ratio of
EW penguin and tree contributions to the isospin amplitude $A_{3/2}$.
In the SU(3) limit, it is calculable in terms of Standard Model
parameters.\cite{us,Fl96} The main uncertainty comes from the present
errors on $|V_{ub}|$ and the mass of the top quark. SU(3)-breaking
effects, which are included in the factorization approximation,
reduce the value of $\delta_{\rm EW}$ by 6\%. In the error analysis
we have assigned a 100\% uncertainty to this estimate. Note that, if
nonfactorizable SU(3) breaking would lead to a further reduction of
$\delta_{\rm EW}$, this would {\em strengthen\/} the bound on $\gamma$
derived below.
The parameter $\varepsilon_{3/2}$ describes the strength of the
$\Delta I=1$ tree amplitude relative to the leading penguin
amplitude. We define a related parameter $\bar\varepsilon_{3/2}$
by writing $\varepsilon_{3/2}=\bar\varepsilon_{3/2}
\sqrt{1-2\varepsilon_a\cos\eta\cos\gamma+\varepsilon_a^2}$, so
that the two quantities agree in the limit $\varepsilon_a\to 0$.
In the SU(3) limit, this new parameter can be determined
experimentally form the relation
\begin{equation}\label{eps}
\bar\varepsilon_{3/2} = \sqrt2\,R_{\rm SU(3)}
\left|\frac{V_{us}}{V_{ud}}\right| \left[
\frac{\mbox{B}(B^\pm\to\pi^\pm\pi^0)}
{\mbox{B}(B^\pm\to\pi^\pm K^0)} \right]^{1/2} \,.
\end{equation}
Factorizable SU(3)-breaking is accounted for by a factor
$R_{\rm SU(3)}\approx f_K/f_\pi\approx 1.2$. This may overestimate
the effect; however, reducing the value of $\bar\varepsilon_{3/2}$
would again {\em strenghten\/} the bound on $\gamma$. We thus feel
comfortable working with a large positive SU(3) correction and
assign an error of 50\% to it to account for nonfactorizable effects.
Using preliminary data reported by the CLEO Collaboration \cite{CLEO}
combined with some theoretical guidance then yields
$\bar\varepsilon_{3/2}=0.24\pm 0.06$.\cite{us} With a precise
measurement of the CP-averaged branching ratios entering (\ref{eps}),
the uncertainty in this number could be reduced significantly.
\section{\boldmath Lower bound on $\gamma$\unboldmath}
Generalizing an idea of Fleischer and Mannel,\cite{FM} a bound on
$\cos\gamma$ can be derived from a measurement of the ratio of the
CP-averaged branching ratios for the two $B^\pm\to\pi K$ decay
modes. The CLEO Collaboration has measured~\cite{CLEO}
\begin{equation}
R_* = \frac{\mbox{B}(B^\pm\to\pi^\pm K^0)}
{2\mbox{B}(B^\pm\to\pi^0 K^\pm)}
= 0.47\pm 0.24 \,.
\end{equation}
It is possible to derive an exact theoretical lower bound on $R_*$
by varying the strong phases $(\eta,\phi)$ independently between $-\pi$
and $\pi$. Knowing the other parameters ($\bar\varepsilon_{3/2}$
from data and $\delta_{\rm EW}$ from theory), this bound implies an
exclusion region for $\cos\gamma$, which becomes larger the smaller
the values of $R_*$ and $\bar\varepsilon_{3/2}$ are. Since the
$B^\pm\to\pi^\pm K^0$ branching ratio enters these two quantities
in an opposite way, it is advantageous to consider the ratio
${\cal R}=(1-\sqrt{R_*})/\bar\varepsilon_{3/2}$, for which the
experimental error on this branching ratio tends to cancel.\cite{Frank}
The current value is ${\cal R}=1.32\pm 0.47$. Theoretically, this ratio
has the advantage of being independent of $\bar\varepsilon_{3/2}$ to
leading order, and one obtains the bound~\cite{us}
\begin{equation}
{\cal R} \le |\delta_{\rm EW}-\cos\gamma|
+ O(\bar\varepsilon_{3/2},\varepsilon_a) \,.
\end{equation}
An exact formula for the higher-order terms can be found in the
literature.\cite{me} For values ${\cal R}>0.8$ the exact bound is
stronger than the approximate one shown above even after the
rescattering effects parametrized by $\varepsilon_a$ are included.
Varying the parameters in the intervals
$0.18\le\bar\varepsilon_{3/2}\le 0.30$ and $0.49\le\delta_{\rm EW}
\le 0.79$, and setting $\varepsilon_a=0.1$, we obtain the bound
shown in Figure~\ref{fig:bound}. Note that the effect of the
rescattering contribution is very small. The table next to the
figure shows the resulting bounds on $|\gamma|$ obtained at different
confidence levels, which we obtained taking into account that in
the Standard Model the largest allowed value of ${\cal R}$ is 1.35.
(This is more conservative than assuming a two-sided Gaussian
distribution.)
\begin{figure}
\epsfxsize=6cm
\epsfbox{bound.eps}
\raisebox{3.2cm}{
\begin{tabular}{|l|c|l|}
\hline
\qquad~\raisebox{0pt}[11pt][4.5pt]{${\cal R}$} & CL & ~~\,Bound \\
\hline
\raisebox{0pt}[11pt][0pt]{1.32} (mean) & \phantom{0}5\% &
$|\gamma|>157^\circ$ \\
\raisebox{0pt}[9pt][0pt]{0.85} ($1\sigma$) & 70\% &
$|\gamma|>93^\circ$ \\
\raisebox{0pt}[9pt][0pt]{0.56} ($1.62\sigma$) & 90\% &
$|\gamma|>71^\circ$ \\
\raisebox{0pt}[9pt][0pt]{0.41} ($1.94\sigma$) & 95\% &
$|\gamma|>60^\circ\,\mbox{or}\!$ \\
\raisebox{0pt}[9pt][6pt]{} & & $|\gamma|<26^\circ$ \\
\hline
\end{tabular}}
\caption{
Left: Theoretical upper bound on the ratio ${\cal R}$ versus
$|\gamma|$ for $\varepsilon_a=0.1$ (solid) and $\varepsilon_a=0$
(dashed). The horizontal line and band show the current experimental
value with its $1\sigma$ variation. Right: Bounds on $|\gamma|$
obtained for specific values of ${\cal R}$.
\label{fig:bound}}
\end{figure}
From a global analysis of the available information on the CKM
matrix elements, one can derive indirect constraints on $\gamma$
that lead to the allowed range $37^\circ\le\gamma\le 118^\circ$,
where the upper limit is determined by the current experimental limit
on $B_s$--$\bar B_s$ mixing, whereas the lower bound is deduced from
the measurement of CP violation in $K$--$\bar K$ mixing.\cite{BaBar}
Without this information, i.e.\ using data from $B$ physics alone,
this lower bound would disappear, and $\gamma=0$ would still be
allowed. The 90\% CL bound on $|\gamma|$ derived here, combined with
the upper bound from $B_s$--$\bar B_s$ mixing, implies that
$71^\circ<|\gamma|<118^\circ$, which is a significant improvement
over the range obtained from the global analysis. This, together with
the fact that $|V_{ub}|\ne 0$ as shown by the existence of semileptonic
$b\to u$ transitions, proves that the unitarity triangle is not
degenerate to a line. In the context of the CKM model, this implies
direct CP violation in $B$ decays.
\section{\boldmath Extraction of $\gamma$\unboldmath}
Ultimately, one would like not only to derive a bound on $\gamma$,
but to measure this parameter directly from a study of CP violation.
Recently, we have described a strategy for extracting $\gamma$ from
$B^\pm\to\pi K$ decays,\cite{us2} which generalizes a method suggested
by Gronau, Rosner and London~\cite{GRL} to include the effects of EW
penguins. This approach has later been generalized to account for
rescattering contributions to the $B^\pm\to\pi^\pm K^0$ decay
amplitudes.\cite{me}
In addition to the ratio $R_*$, one considers the following
combination of the direct CP asymmetries in the decays
$B^\pm\to\pi K$:
\begin{equation}
\widetilde A \equiv \frac{A_{\rm CP}(\pi^0 K^\pm)}{R_*}
- A_{\rm CP}(\pi^\pm K^0) \,.
\end{equation}
With this definition, the rescattering effects parametrized by
$\varepsilon_a$ are suppressed by a factor of $\bar\varepsilon_{3/2}$
and thus reduced to the percent level. The same is true for the ratio
$R_*$. Explicitly, we have
\begin{eqnarray}
R_*^{-1} &=& 1 + 2\bar\varepsilon_{3/2}\cos\phi\,
(\delta_{\rm EW}-\cos\gamma) \nonumber\\
&&\mbox{}+ \bar\varepsilon_{3/2}^2\,
(1-2\delta_{\rm EW}\cos\gamma+\delta_{\rm EW}^2)
+ O(\bar\varepsilon_{3/2}\,\varepsilon_a) \,, \nonumber\\
\widetilde A &=& 2\bar\varepsilon_{3/2} \sin\gamma \sin\phi
+ O(\bar\varepsilon_{3/2}\,\varepsilon_a) \,.
\end{eqnarray}
For fixed values of $\bar\varepsilon_{3/2}$ and $\delta_{\rm EW}$,
these equations define contours in the $(\gamma,\phi)$ plane. When
the rescattering corrections from $\varepsilon_a$ are included,
these contours become narrow bands. From the intersection of the
contour bands for $R_*$ and $\widetilde A$ one obtains the values of
$\gamma$ and the strong phase $\phi$ up to possible discrete
ambiguities. For some typical numerical examples, the theoretical
uncertainties on the extracted values of $\gamma$ resulting from the
variation of the input parameters $\bar\varepsilon_{3/2}$,
$\delta_{\rm EW}$ and $\varepsilon_a$ are found to add up to a
total error of order $\delta\gamma\sim 10^\circ$.\cite{us2,me} A
precise determination of this error requires, however, to know the
actual values of $R_*$ and $\widetilde A$. For instance, Gronau
and Pirjol~\cite{Pirj} find a larger error for the specific case
where the product $|\sin\gamma\sin\phi|$ is {\em very\/} close to 1,
which would imply a value of the CP asymmetry $\widetilde A$ close
to 50\%.
\section{Conclusions}
Measurements of the exclusive hadronic decays $B\to\pi K$ provide
interesting information on the weak phase $\gamma$. Using CLEO data
on the CP-averaged branching ratios of the two charged decay modes,
we have derived the bound $|\gamma|>71^\circ$ at 90\% CL. This bound
is stronger than the lower bound derived from the global analysis
of all other information on the CKM matrix. Combined with constraints
from $B_s$--$\bar B_s$ mixing and semileptonic $b\to u$ decays, and
in the context of the CKM model, it proves the existence of direct
CP violation in $B$ decays.
\newpage
\section*{Acknowledgments}
It is a pleasure to thank the organizers of WIN99, C.A.~Dominguez,
R.D. Viollier and their staff, for arranging a marvellous meeting in
a fantastic setting. I am grateful to Jon Rosner for collaboration
on most of the work reported here. I also wish to thank Frank
W\"urthwein for useful comments. This work was supported by the
Department of Energy under contract DE--AC03--76SF00515.
|
\section{Introduction}
\label{sec:intro}
The detection of fluctuations in the cosmic microwave background (CMB)
by the Differential Microwave Radiometer (DMR) on board the Cosmic
Microwave Explorer (COBE) satellite (Smoot {\frenchspacing\it et al.}\ 1992) began a new
era for studies of the Universe on large scales. Current and planned
experiments offer the promise of tight constraints on cosmological
parameters, including both quantities important for observational
astronomy and inflationary parameters which may provide the ultimate
testing ground for fundamental particle physics, relating phenomena on
the largest and smallest observable scales.
A key cosmological constraint from observations of the primordial
density field is its general statistical nature. In most inflationary
scenarios, the density is a Gaussian random field, (although see, {\frenchspacing\it e.g.},
Peebles 1999). This implies that the joint probability distribution
of the $N\sim 4000$ temperatures in the galaxy-cut DMR sky map is a
multivariate Gaussian. In contrast, topological defect models predict
a non-Gaussian density field ({\frenchspacing\it e.g.}, Avelino {{\frenchspacing\it et al.}} 1998). However, an
analysis by Pen {{\frenchspacing\it et al.}} (1997) suggests that defects are inconsistent
with the observed power spectrum of fluctuations. At present,
inflationary models, or at least generically Gaussian models, dominate
the literature on large-scale structure.
It is therefore quite intriguing that two groups, Ferreira, Magueijo
\& Gorski (1998, hereafter FMG) and Pando, Valls-Gabaud \& Fang (1998,
hereafter PVF), claim that the CMB fluctuations measured by COBE DMR
are non-Gaussian. If substantiated, these results could potentially
rule out standard inflation as the primary mechanism for cosmic
structure formation in the early Universe. The results are all the
more surprising because the literature documents considerable previous
effort to identify non-Gaussianity in the CMB (see references given in
\S\ref{DebunkSec}), all of which failed.
In this {\em Letter}, we revisit the question of whether COBE DMR
rules out Gaussianity. We begin with a series of new tests of
Gaussianity based on eigenmode analyses of the DMR map
(\S\ref{EigenSec}). We then consider the results of FMG and PVF in
\sec{DebunkSec}, exploring their sensitivity to individual modes and
spatial features as well as the significance of these results in light
of the other statistical tests which have been applied to the
data. We present our conclusions in \sec{ConclusionSec}.
\section{Statistics of eigenmodes}\label{EigenSec}
We analyze a sky map formed of the combined 53 and 90 GHz four year DMR
data pixelized at resolution 6 in Galactic coordinates, with the
``custom'' Galaxy sky mask (Bennett {{\frenchspacing\it et al.}} 1996), and with monopole
and dipole contributions removed as in Tegmark \& Bunn (1995). The
resulting data set consists of temperatures for 3,881 pixels in the
sky, which we arrange in a vector ${\bf x}$. We take the covariance matrix
${\bf C}\equiv \expec{{\bf x}\x^t}$ to be of the form ${\bf C}={\bf S}+{\bf N}$, where the
noise matrix ${\bf N}$ is diagonal and the signal matrix ${\bf S}$ is a
Harrison-Zel'dovich power spectrum normalized to $Q_{rms-PS} = 18 \ \mu$k
({\frenchspacing\it e.g.}, Bennett {{\frenchspacing\it et al.}} 1996).
The pixelized data are both correlated and noisy, hence we subject the
map ${\bf x}$ to both a principal component analysis (PCA) and a
signal-to-noise eigenmode analysis (SNA), two standard astrophysical
tools. Both of these procedures involve expanding the data in a new
basis, the eigenvectors of ${\bf C}$ for the PCA case and the eigenvectors
of ${\bf N}^{-1/2}{\bf S}{\bf N}^{-1/2}$ for the SNA case. The eigenvectors are
sorted by decreasing eigenvalue and normalized so that the expansion
coefficients have unit variance. For the SNA (Bond 1994; Bunn \&
Sugiyama 1995; Tegmark {{\frenchspacing\it et al.}} 1997), the modes are listed in order of
decreasing signal-to-noise level. For the PCA, the modes explain
successively less and less of the variance in the data. Since the DMR
noise per pixel does not fluctuate much, the two methods give similar
results. The first few hundred modes contain essentially all the
cosmological information, and probe successively smaller angular
scales (Bond 1994; Bunn \& Sugiyama 1995; Bunn \& White 1996). We use
the top 250 modes for the Gaussianity tests described below.
The purpose of this exercise is two-fold: First, we can determine how
many cosmologically significant degrees of freedom a given statistical
test should consider. Second, the decomposition into uncorrelated
eigenmodes allows the data to be cast as a list of random numbers
which, under the null hypothesis that the DMR data are
Gaussian\footnote{Note that as long as the true CMB sky is Gaussian,
our data set will be Gaussian as well: Both the smoothing done by the
DMR beam, our galaxy cut and our monopole and dipole removal are
linear operations, and all linear operations preserve Gaussianity.},
will be independent and normally distributed. Although
the statistics of these samples may not be completely testable in
practice, we can still constrain general properties of the COBE data.
We run both the lists of 250 PCA and SNA entries and the entire list
of 3877 numbers (the rank of the covariance matrix after monopole and
dipole subtraction) from the PCA basis through a smattering of tests,
first for Gaussianity of the individual list elements. The null
hypothesis cleanly passes Kolmogorov-Smirnoff and $\chi$-square tests,
along with tests of cumulants up to fourth order and of the
significance of the top few outliers. None of these tests manage to
reject the Gaussian null hypothesis with 95\% confidence. Note that
these tests are sensitive only to the 1-point distribution of mode
amplitudes, not to correlations between modes. This is strong though
not irrefutable evidence that if the DMR data are non-Gaussian then
mode correlations, not mode amplitudes, are responsible.
The next step in testing the Gaussian hypothesis is to look for mode
correlations. This is a difficult thing to do in any exhaustive way,
even for our short lists of 250 elements, and a thorough treatment of
this problem is beyond the scope of this {\em Letter}. We note only
that no correlations were detected above the 95\% confidence level in
tests of second, third and fourth order $N$-point correlations. We
also used mode amplitudes to simulate rolls of a die and examine the
one- and two-point distributions of outcomes to see if the die is
loaded. If one wishes to play dice with the Universe, evidently it
would be a fair game.
\begin{figure*}[tb]
\centerline{\epsfxsize=12cm\epsffile{zoom.eps}}
\figcaption{
The galaxy-cut COBE DMR map is shown in Aitoff projection.
The inset shows the five pixels whose omission makes the
FMG effect go away, with the ellipse indicating the
FWHM size and the projected shape of the DMR beam
in that region.
}
\end{figure*}
\section{Reports of Non-Gaussianity}
\label{DebunkSec}
The above tests of the COBE data (26 in all) join a sizable list of
previous results without significant detections of non-Gaussianity.
For the 53 GHz DMR 1 year data, the three-point function was studied
by Luo (1994, one test) and Hinshaw {{\frenchspacing\it et al.}} (1994, two tests) while
Smoot {{\frenchspacing\it et al.}} (1994) considered the topological genus and kurtosis
(two tests). For the 53 and 90 GHz DMR 2 year data, Hinshaw {{\frenchspacing\it et al.}}
(1995) studied the equilateral and pseudocollapsed three-point
function at three $\ell$-cuts with 12 tests in total; the most extreme
gave a 98\% non-Gaussianity detection, but was deemed to suffer from a
known noise problem. Kogut {\frenchspacing\it et al.}\ (1996) tested the DMR 4 year data
for three-point correlations, genus and peak correlations (4 tests in
all), while Heavens (1998) analyzed this same data set using an
optimized bispectrum statistic on 5 different scales. Gazta\~naga
{{\frenchspacing\it et al.}} (1998) performed 5 variance-of-variance tests with the
strongest rejection of Gaussianity being at the 91\% level. Most
recently, Diego {{\frenchspacing\it et al.}} (1999) concluded that the DMR data were
consistent with Gaussianity in a partition function analysis.
On the other hand, there are two Gaussianity tests which the data
reportedly fail. One is based on the bispectrum statistic of FMG (at
98\% confidence) and other on the fourth-order wavelet statistic
proposed by PVF (at 99\% confidence). These are strong signals of
non-Gaussianity, but some caution is in order. Given that the DMR
data have been subjected to the 32 other published tests cited above
which provide no evidence of non-Gaussianity (not to mention the 26
reported here and any tests, including some of our own, which were not
reported because they yielded null results), are the FMG and PVF
results simply expected outliers in the distribution of test results?
To address this point, suppose we try to rule out some null hypothesis
by subjecting a data set to $n$ independent statistical tests, and
that the most successful one rules it out at a confidence level $p$,
say 99\%. How significant is this really? Let $p_i$ denote the
confidence level obtained from the $i^{th}$ test, with
$p_{\rm max}\equiv\max\{p_i\}$ corresponding to the most successful test.
The probability of getting a less extreme result is then
\beq{DilutionEq}
P(p_{\rm max}<p) = \prod_{i=1}^n P(p_i<p)
= \prod_{i=1}^n p
= p^n \, .
\end{equation}
For example, the most extreme S/N eigenmode coefficient in
\sec{EigenSec} is a 3.3-$\sigma$ outlier. If that one coefficient was
all we had, then we would reject Gaussianity at the 99.9\% level.
However, we have 250 independent numbers and \eq{DilutionEq} shows
that our level of confidence in rejection from that one extreme
coefficient is only $0.999^{250}\approx 78\%$. Similarly, the list of
34 published Gaussianity tests mentioned above contains one which
rules out the null hypothesis at the 99\% level. If these tests were
independent, we could reject Gaussianity with only $0.99^{34}\approx
71\%$ confidence. Of course the tests are not strictly independent.
Yet if some of them capture only subsets of the information contained
in the $\sim 4000$ COBE data points, then they may be effectively
independent of each other. With this in mind we examine the FMG and
PVF tests in more detail.
\subsection{The wavelet test}
The detection reported by PVF of a non-Gaussian signal is made with a
measure of fourth-order correlations between wavelet coefficients of
the DMR data. The coefficients are obtained using a discrete transform
of the northern sixth of the sky map, after the spherical plane of the
sky has been projected onto the face of a cube. Taken alone, this
measure reportedly gives a detection of non-Gaussian signal at 99\%
confidence. But we have more information, even about PVF's wavelets: a
second projection of the sky map onto the opposite face of the cube
(i.e., the opposite hemisphere of the sky) lies roughly at the 40\%
confidence level. Together, a joint two-faced wavelet analysis gives a
weaker rejection of the Gaussian hypothesis, formally at 97\%
confidence. Furthermore, the PVF detection is claimed only for
wavelets on one specific scale even though they seek similar
detections on two other scales but do not find them. Likewise, a
detection was sought but not found for third-order moments. With
$n=2\times 3\times 2$, \eq{DilutionEq} predicts a much lower
confidence level, $0.99^{12}\approx 89\%$, for the claimed detection.
In an analysis based on compact but smooth wavelets (e.g., Bromley
1994), we confirm the existence of a strong (99.6\%) non-Gaussian
outlier at the 11$^\circ$+22$^\circ$ scales reported by PVF.
Interestingly, we can make the entire non-Gaussian signal vanish by
simply zeroing or flipping the sign of a single, modestly rare
principal component amplitude (a 2.7-$\sigma$ fluctuation of the
90$^{\rm th}$ mode). Also the 17$^{\rm th}$ eigenmodes from both the
PCA and SNA strongly affect the wavelet detection. In the S/N basis,
this mode has an amplitude of 2.1-$\sigma$ (slightly less in the
principal component vector); by zeroing the amplitude of these modes,
the wavelet statistic yields less than a 2-$\sigma$ detection.
Furthermore, when set to zero, a single spot in the sky the size of
the COBE beam (centered at $b = 64.6^\circ$, $l = 40.6^\circ$), also
cuts the non-Gaussian signal down to a similar level. Of the six
pixels in this spot, three are within one standard deviation of the
expected noise, while two are at the 2.6-$\sigma_n$ level and the
third is a strong outlier at 3.1-$\sigma_n$. (How this spot affects
the PVF results depends on unspecified details of their analysis.)
\subsection{The bispectrum test}
FMG introduce a measure, $I^3_{\ell}$, based on averaged triplets of
projection coefficients from even-multipole spherical harmonics.
Non-Gaussian behavior is seen only at $\ell = 16$, but FMG are careful
to consider the fact that the bispectrum at eight other $\ell$-values
are individually consistent with the Gaussian hypothesis. The reported
confidence of the non-Gaussian detection is 98\%.
There is nonetheless a possibility that the nine $I^3_{\ell}$\ values given by
FMG are sensitive to only a fraction of the information in the COBE
data. Although there is some spherical harmonic mode coupling as a
result of the sky mask, the odd $\ell$ multipoles are largely missing as
well as multipoles above $\ell = 18$. There may also be dependence on
localized noise. We emphasize this latter point by setting to zero the
five pixels in the beam-size spot on the sky centered at Galactic
latitude $b = 39.5^\circ$ and longitude $\ell = 257^\circ$, shown in
Fig.~1. The value of $I^3_{16}$\ falls from about 0.92 to 0.78,
approximately a 98\% detection on its own\footnote{Note that the
effect of zeroing the spot depends somewhat on details of removing
monopole and dipole contributions to the DMR maps. If explicit
removal (e.g., Tegmark \& Bunn 1995) is not performed on the custom
cut map, as is apparently the case in the bispectrum analysis of
Magueijo, Ferreira \& Gorski (1999; Fig.~1 therein), then the effect
of removing the spot (actually a nearest-neighbor) is to lower
$I^3_{16}$\ to 0.66.}, but well below 2-$\sigma$ when taken in
conjunction with the other eight $I^3_{\ell}$\ values shown in Fig.~2
($.98^9\approx 84\%$ if the 9 values were uncorrelated). Note that
there is a single, rare pixel brightness value in the spot. In
units of the expected noise fluctuations, it is at the level of
$-3.5\sigma_n$, and zeroing it alone cuts $I^3_{16}$\ by 10\%.
\begin{figure}[tb]
\vskip-2.7cm
\hglue-0.8cm
\centerline{\hglue1cm{\vbox{\epsfxsize=10.5cm\epsfbox{figell.ps}}}}
\vskip-2.3cm
\figcaption{
The bispectrum statistic of FMG for various multipoles $\ell$.
The dark circles are the full galaxy-cut COBE data (cf. Fig.~1 in FMG), and
the open circles are the data after zeroing 5 pixels inside a single
COBE beam centered at $b = 39.5^\circ$, $l = 257^\circ$). The
triangles result when the amplitude of mode 151 (containing 0.15\% of
the total power) is set to zero.
}
\label{fig:bis}
\end{figure}
The bispectrum statistic is also highly sensitive to individual
eigenmodes. Zeroing or flipping the sign of principal component 151
causes $I^3_{16}$\ to drop from 0.92 to the unambiguously Gaussian values
of 0.61 and 0.20, respectively.
Furthermore, there is sensitivity to S/N eigenmode 224, a 3-$\sigma$
fluctuation (the second most extreme of the first 250 modes) which is
more strongly coupled to noise than cosmic structure (its S/N
eigenvalue is 0.4). Zeroing this knocks $I^3_{16}$\ to 0.74; flipping the
sign causes the value to fall to 0.52, a decidedly Gaussian value.
Since both of these eigenmodes are dominated by noise rather than
cosmic signal, it is possible that the source of the alleged
non-Gaussianity is detector noise rather than CMB.
Note that in the above analyses, we systematically searched for the
modes or pixels which affected the wavelet and bispectrum statistics
the most. In each case we found only a few which cause more than an
insignificant (several percent) change in these measures, and
interestingly the significant modes or pixels were different for the
two measures.
Using randomly generated Gaussian skymaps selected for
apparent non-Gaussianity similar to the DMR data, we also checked that
it is quite common for a Gaussian map to give a wavelet or bispectrum
measure that is sensitive to only a few individual pixels or modes,
just as in the DMR case.
\section{Conclusion}
\label{ConclusionSec}
The problem of the statistical nature of the CMB may be cast in a
conceptually simple form: just use the observed temperature
fluctuations to make a random number generator, based on assumed
statistical properties, and test its quality. Here we have used the
53+90 GHz COBE sky map to generate lists of putative random numbers
with principal components and signal-to-noise eigenmodes. In both
cases the one-point distribution is manifestly Gaussian. If the CMB
on COBE scales is non-Gaussian, it is the result of correlations
between modes. Unfortunately, exhaustive tests of mode correlations
are not feasible. A few tests which can pick up a fair range of
non-Gaussian behavior were performed and no evidence of mode
correlations was found.
Here we have also considered the two statistics which reportedly detect
non-Gaussianity in the COBE data. Both detections turn out to be
fragile in the sense that they vanish when a single DMR beam spot or
a single eigenmode is removed. Moreover, we found that the detection
by PVF, based on wavelets alone, was less significant than originally
claimed. Even so, with the dozens of different Gaussianity tests that
have now been published, it would not be surprising if a perfectly
valid analysis rejected Gaussianity at say 98\% confidence purely by
accident.
Our results cast some doubt on the significance of the claimed
non-Gaussian behavior in the CMB. If the reported detections are real
nonetheless, then the eigenmodes and COBE-beam spots that we isolated
for the wavelet and bispectrum statistics are candidates for potential
non-Gaussian sources in the CMB. This latter possibility would perhaps
be more satisfying if both measures were coupling to the same
non-Gaussian structure in the sky. However, this is not obviously the
case, since both the bispectrum statistic and the wavelet measure show
virtually no sensitivity to the sky spots and eigenmodes which so
dramatically affect the other.
It is generally much easier to show that a bad random number generator
is bad then to prove that a good one is good. Indeed, the results
reported here fail to demonstrate that the CMB really is
Gaussian. Conversely, the search for non-Gaussianity is also something
of an uphill battle, a fight against the central limit theorem which
causes both instrumental effects and the linear combinations involved
in the eigenmode expansions to make things look more Gaussian.
Therefore statistical measures should be tuned for the specific type
of non-Gaussianity that physical models predict. This approach is
taken in many recent studies ({\frenchspacing\it e.g.}, Cayon \& Smoot 1995; Magueijo 1995;
Torres {{\frenchspacing\it et al.}} 1995; Gangui 1996; Gangui \& Mollerach 1996; Ferreira
\& Magueijo 1997; Ferreira {{\frenchspacing\it et al.}} 1997; Barrieiro {{\frenchspacing\it et al.}} 1998, Lewin
{{\frenchspacing\it et al.}} 1999; Popa 1998) with an eye toward upcoming, high-resolution
CMB data.
We thank Ang\'elica de Oliveira-Costa, Al Kogut, Alex Lewin, Bill
Press, George Rybicki, and Nelson Beebe for useful comments.
BCB acknowledges partial support from NSF Grant PHY 95-07695 and the use
of supercomputing resources provided by NASA/JPL and Caltech/CACR. MT
was funded by NASA though grant NAG5-6034 and Hubble Fellowship
HF-01084.01-96A from STScI, operated by AURA, {\frenchspacing Inc.}
under NASA contract NAS5-26555.
\vskip-1cm
|
\section{Introduction}
It is now almost universally accepted that any generic black
hole\footnote{Supersymmetric black holes, i.e., those associated with a
covariantly conserved Killing spinor, are usually extremal and do not emit
Hawking radiation. Such special black holes are not considered here.} with
infalling quantum matter fields radiates like a
black body in equilibrium
at a temperature $T_{bh} \sim \kappa/2 \pi$, where, $\kappa$ is the
surface gravity at the horizon of the black hole \cite{hawk}.
Consequently, infalling bose fields are radiated out in a thermal
Bose-Einstein (Planckian) distribution (modulo some `grey-body' factors)
while fermions are radiated out in a Fermi-Dirac distribution. On the
basis of this alone, one might expect that the supersymmetry manifest at
past
null infinity in a system of bosons and fermions will not survive such
radiation. This expectation seems to be borne out in detail in an
investigation performed two years ago \cite{pm}, in which the standard
criterion of spontaneous supersymmetry breakdown is used, namely the
non-vanishing (or otherwise) of the vacuum expectation value of the
supersymmetry variation
of a fermionic operator at future null infinity. The evaluation of
the vacuum expectation value (vev) in ref. \cite{pm} follows
Hawking's original approach involving a zero temperature quantum field
theory in the black hole background, and is generic in nature. However,
in that derivation, as elsewhere, one has tended to ignore the
fact that the spinorial supersymmetry parameter is actually an element of
a Grassmann algebra (an $a$-number) rather than a $c$-number. In other
words, unlike a $c$-number parameter which commutes with all
operators of the theory, the supersymmetry parameter {\it anti}commutes
with fermionic operators instead. Clearly, this may have serious
implications for evaluation of Green's functions and the like involving
strings of fermionic and bosonic operators, such as the issue of
supersymmetry breaking entails. Our concern here is with possible
ramifications for black hole radiance. Recall that the phenomenon of
black hole radiance is based upon particle creation in the
gravitational field of a black hole. Thus, evaluation of vevs of operators
defined at future null infinity will
involve matrix elements of such operators between states (at future null
infinity) of non-zero fermion number. It is here that a na\"ive handling
of the supersymmetry parameter is most likely to differ from a careful
one. If sharp disparities arise, the conventional wisdom that black hole
radiance breaks supersymmetry, is bound to be challenged.
Section II of the paper is a brief recapitulation of the main tenets of
the earlier work. In section III, we attempt a more careful evaluation of
the relevant vacuum expectation value, to see if supersymmetry can indeed
be preserved in black hole radiance. In section IV, a comparative analysis
of the supersymmetric model in flat spacetime in presence of a heat bath
at a finite temperature is presented. Our conclusions and outlook are
presented in the final section.
\section{The earlier formulation surveyed}
We focus on a situation where, at past null infinity (${\cal I}^-$), there
exists a globally supersymmetric model of noninteracting massless complex
scalar and chiral spinor fields. Now, any state on ${\cal
I}^-$ will evolve into a state on the event horizon (${\cal H}^+$),
belonging to one of two mutually exclusive (in absence of backreaction)
classes, viz., those which are purely outgoing, i.e., have zero Cauchy
data on ${\cal H}^+$ and support on ${\cal I}^+$, and those which have
zero Cauchy data on ${\cal I}^+$ and support on ${\cal H}^+$. As is
well-known \cite{hawk}, an inherent ambiguity in the latter is chiefly
responsible for the thermalization of the radiation received at ${\cal
I}^+$.
The scalar and spinor fields in our model have the following expansion
(at ${\cal I}^-$),
\begin{eqnarray}
{\phi} ~ &=& ~\sum_k \frac {1}{\sqrt{2 \omega_k}}~\left ( a^B_k f^B_k~+~
b_k^{B ~ \dag}
{\bar f_k^B} \right) ~\nonumber \\
{\psi_+} ~ &=& ~\sum_k \frac {1}{\sqrt{2 \omega_k}}~ \left( a^F_{k,+}
f^F_k~+~b_{k,-}^{F~ \dag} {\bar f_k^F} \right) u_{k,+}~, \label{flxp}
\end{eqnarray}
where, the $\{ f_k \}$ are complete orthonormal sets of solutions of the
respective field
equations, with positive frequencies only at ${\cal I}^-$, and $u_{k,+}$
is a positively
chiral spinor, reflecting the chirality of
$\psi_+$. The creation-annihilation operators obey the usual algebra,
with $B$ ($F$)
signifying Bose (Fermi). The conserved N\"other supersymmetry charge is
given in terms of
these creation-annihilation operators by (at ${\cal I}^-$)
\begin{equation}
Q_+({\cal I}^-)~=~\sum_k \left( a^F_{k,+}
{b^B_k}^{\dag}~-~{b^F_{k,-}}^{\dag} a^B_k \right) u_+(k)~,\label{qu}
\end{equation}
and annihilates the vacuum state $|0_- \rangle$ defined by
\begin{equation}
a^{B,F}_k |0_-\rangle~=~0~=~b^{B,F}_k |0_- \rangle~~. \end{equation}
The existence of two disjoint classes of states at the horizon, as
mentioned earlier, implies that the fields also admit the expansion
\cite{hawk}
\begin{eqnarray}
\phi ~ &=& ~ \sum_k \frac{1}{ \sqrt{2
\omega_k}} ~ \left(~ A_k^B p_k^B~+~B_k^{B~ \dag} {\bar p}_k^B ~+~A_k^{'B}
q_k^B~+~ B_k^{'B~ \dag} {\bar q_k^B} ~ \right)~\nonumber \\
\psi_+ ~ &=&
~\sum_k \frac{1}{\sqrt{\omega_k}}~ \left( A_{k,+}^F p_k^F~+~ B_{k,+}^{F~
\dag} {\bar p}_k^B ~+~A_{k,+}^{'F} q_k^F~+~ B_{k,-}^{'F~ \dag} {\bar
q_k^F} \right)~u_+(k)~~, \label{flxp+}
\end{eqnarray}
where, $\{p_k\}$ are
purely outgoing orthonormal sets of solutions of the respective field
equations with positive frequencies at ${\cal I}^+$, while $\{q_k\}$ are
orthonormal sets of solutions with no outgoing component. The final vacuum
state $|0_+ \rangle$, defined by the requirement
\begin{equation}
A^{B,F}|0_+ \rangle~=~0~=B^{B,F} |0_+ \rangle ~=~A'^{B,F}
|0_+ \rangle~=~B'^{B,F} |0_+ \rangle~
\end{equation}
is not unique, because of the inherent ambiguity in defining positive
frequency for the $\{q_k\}$; in fact, one can write $|0_+\rangle = |0_I
\rangle |0_H \rangle$
with the unprimed (primed) operators acting on $|0_I>$ ($|0_H>$). A
supersymmetry charge
$Q({\cal I}^+)$ may indeed be defined, analogously to eqn. (\ref{qu}), in
terms of the unprimed operators, and that $Q({\cal I}^+) |0_+
\rangle~=~0$. Such
a charge also satisfies the $N=1$ superalgebra at ${\cal I}^+$.
The field operators $a_k, b_k$ at ${\cal I}^-$ are of course related to the $A_k, B_k$
and $A'_k, B'_k$ through the Bogoliubov transformations
\begin{eqnarray}
A_k^B~&=&~ \sum_{k'} \left(~\alpha_{kk'}^B a_{k'}^B~+~\beta_{kk'}^B
b_{k'}^{B ~\dag}~\right)~\nonumber\\
B_k^B~& =&~ \sum_{k'} \left(~\alpha_{kk'}^B b_{k'}^B ~+~\beta_{kk'}^B
a_{k'}^{B~\dag}~
\right)~\nonumber \\
A_{k,+}^ F~&=&~\sum_{k'} \left(~\alpha^F_{kk'} a^F_{k', +}~+~\beta^F_{kk'}
b_{k',
-}^{F~ \dag}~\right)~ \nonumber \\
B^F_{k',- }~&=&~\sum_{k'} \left(~\alpha_{kk'}^F b_{k',-}^F~+~\beta_{kk'}^F
a^{F~ \dag}_{k',
+}~\right)~, \label{bog}
\end{eqnarray}
and similarly for the primed operators. We notice in passing that
\begin{equation}
Q({\cal I}^-) |0_+ \rangle~\neq~ 0~,~ Q({\cal I}^+) |0_-
\rangle~\neq~0~,~ for
~\beta^{B,F}~\neq~0~.\label{hnt} \end{equation}
The issue that we now wish to focus on is whether the radiated system of
particles has $N=1$ spacetime supersymmetry. To address this question,
recall that vacuum
expectation values (vevs) of observables at future null infinity are
defined by \cite{hawk}
\begin{equation}
\langle~ {\cal O}~ \rangle~\equiv~\langle 0_-|~ {\cal O}~ |0_-
\rangle~=~Tr~
(\rho ~{\cal O})~ \label{vev}
\end{equation}
where, $\rho$ is the density operator. The trace essentially averages over the
(nonunique) states going through the horizon, thus rendering the vevs of
observables (at
${\cal I}^+$) free of ambiguities. We also recall that a
sufficient condition for spontaneous supersymmetry breaking is the
existence of {\it a}
fermionic operator ${\cal O}$ which, upon a supertransformation, yields
an operator with
non-vanishing vev, i.e., $ \langle~ \delta_S {\cal O}~ \rangle~\neq~0$.
Thus, if one is
able to show that
for {\it all} fermionic observables ${\cal O}({\cal I}^+)$,
\begin{equation}
\langle~ \delta_S {\cal O}({\cal I}^+)~ \rangle~=~0~, \label{ss}
\end{equation}
then we are guaranteed that the outgoing particles form a supermultiplet.
However, this is not the case, as is not difficult to see; for, consider
the supercharge
operator itself at ${\cal I}^+$. Using the supersymmetry algebra, it can
be shown that
\begin{equation}
\langle~ \delta_S {\bar Q}_{\dot \alpha}({\cal I}^+)~
\rangle~=~\epsilon^{\beta}~ \langle~ P_{\beta {\dot \alpha}} ({\cal
I}^+)~ \rangle~,\label{mom} \end{equation}
where $P_{\beta {\dot \alpha}}$ is the momentum operator of the theory
and $\epsilon^{\beta}$ the spinorial supersymmetry parameter.
In our free field theory, the rhs
of (\ref{mom}) is trivial to calculate, using eq. (\ref{bog}) above, so
that we obtain, suppressing obvious indices,
\begin{eqnarray}
\langle~\delta_S {\bar Q}({\cal I}^+) ~\rangle~ &=& \epsilon \sum_k k ~
\langle~N^B_k~+~N^F_k~ \rangle~\nonumber \\
&=& \epsilon \sum_{k,k'} k \left(~
|\beta_{kk'}^B|^2~+~|\beta^F_{kk'}|^2 ~\right)~~.\label{ssbr}
\end{eqnarray}
Thus, supersymmetry is seemingly spontaneously broken in the sense
described above, so long as the Bogoliubov coefficients $\beta^{B,F}$ are
non-vanishing In fact, we know from Hawking's seminal work \cite{hawk}
that
\begin{eqnarray}
\langle ~N_k^B~ \rangle~ &=& ~ \sum_{k'}
|\beta^B_{kk'}|^2~=~|t^B_k|^2~\left(~e^{2\pi
|k|/\kappa}~-~1~\right)^{-1}~\nonumber \\
\langle~N_k^F~\rangle~ &=&~ \sum_{k'}
|\beta^F_{kk'}|^2~=~|t^F_k|^2~\left(~e^{2\pi|k|/
\kappa}~+~1~\right)^{-1}~.\label{haw}
\end{eqnarray}
Here, $\kappa$ is the surface gravity of the black hole, and
$|t_k^{B,F}|^2$ the transmission coefficients through the potential barrier of the
black hole for bosons, fermions respectively.
\section{A more careful evaluation}
Using the definition (\ref{vev}) above, the aim is
to evaluate the trace $Tr~\left(~ \rho~[~\epsilon~ Q~,~ {\bar
Q}~]~\right) $. Treating $\epsilon~Q $ as a {\it bosonic} operator and
using the cyclicity of traces, it is easy to see that,
\begin{equation}
\langle~ \delta_S {\cal O}~\rangle~=~Tr~\left(~ [~{\bar
Q}~,~\rho~]~\epsilon~Q~\right)~. \label{trc}
\end{equation}
Thus, the issue of supersymmetry breakdown now depends crucially on
whether the density operator $\rho$ commutes with the supersymmetry
generator ${\bar Q}$ at ${\cal I}^+$. As an operator relation this is
not obvious since we do not know the density operator as a function of
the basic field operators $A^{B,F}~,~B^{B,F}$. In Hawking's approach,
one can only unambiguously determine the diagonal elements of the
density matrix. One would expect the determination of such elements to
be enough to ascertain whether the vev $\langle~\delta_S {\bar
Q}~\rangle$ is non-vanishing.
The basic point of departure from earlier approaches (including ours) is
the property that for any fermionic
operator ${\cal O}~,~ \epsilon {\cal O} = - {\cal O} \epsilon$. That is
to say that $\{ \epsilon, A^F \} = 0$ and similarly for $B^F$. It
follows that, for normalized states (at ${\cal I}^+$) with $n^F_k$
fermions with momentum $k$, we must have
\begin{equation}
\langle n^F_k|~ \epsilon~ | n^F_k \rangle ~=~(-)^{n^F_k} \epsilon~.
\label{enef}
\end{equation}
In our earlier approach \cite{pm}, the {\it rhs} of eq. (\ref{enef})
would not have contained the first factor. This does have an immediate
import for our calculation of $\langle~\delta_S {\cal O}~\rangle $ above
in eq. (\ref{mom}). Rather than expanding the commutator in the
variation $\delta_S {\bar Q}$ as done above, we follow our earlier
step eq. (\ref{vev}) of using the supersymmetry algebra and rewriting
(\ref{vev}) as given in eq. (\ref{mom}),
\begin{equation}
\langle~ \delta_S {\bar Q}_{\dot \alpha}({\cal I}^+)~
\rangle~=~Tr ~\left(~\epsilon^{\beta}~ ~ P_{\beta {\dot \alpha}} ({\cal
I}^+)~ \rho~\right ).\label{mom1}
\end{equation}
The Hilbert space of this non-interacting theory ${\cal H} \sim {\cal H}_B
\otimes {\cal H}_F$ so that, changing to the occupation number basis of
the infinite system of uncoupled bose and fermi oscillators, labelled by
momentum $k_{\alpha {\dot \beta}}$, the states are expressed as
$|n^B_k~,~n^F_k~\rangle $. Assuming without loss of generality, a discrete
momentum spectrum, eq. (\ref{mom1}) may be expressed as
\begin{equation}
\langle~ \delta_S {\bar Q}_{\dot \alpha}({\cal I}^+)~
\rangle~=~\epsilon^{\beta}~\sum_k k_{\beta
{\dot \alpha}}~ \sum_{n^B_k, n^F_k}~\rho_{n^B_k n^F_k,n^B_k
n^F_k}~(-)^{n^F_k}~\left(~n^B_k~+~n^F_k~\right)~. \label{rho}
\end{equation}
Since the relevant density matrix elements are
uniquely determined by states at future infinity (the horizon states being
averaged over) where the system of particles are still non-interacting,
the density operator can be factorized as $\rho = \rho^B ~\rho^F$ where
$\rho^B (\rho^F)$ acts on bosonic (fermionic) states $|n^B_k \rangle$
($|n^F_k \rangle $) alone respectively.
Recall now the elementary fact that for a given momentum $k$,
$n^F_k~=~0~,~1$. Using eq. (\ref{enef}), (\ref{mom1}) immediately yields
\begin{equation}
\langle~ \delta_S {\bar Q}_{\dot \alpha}~\rangle~=~\epsilon^{\beta}~
\sum_k k_{\beta
{\dot \alpha}}~ \sum_{n^B_k=0}^{\infty}~n^B_k~\left(
~\rho^B_{n^B_k,n^B_k}~\rho^F_{00}~-~\rho^B_{n^B_k-1,n^B_k-1}~\rho^F_{11}
~\right)~~. \label{dens}
\end{equation}
Not surprisingly, bose-fermi pairing, characteristic of supersymmetric
theories, seems to appear here as well. Thus, if the density operator does
indeed commute with the supersymmetry generator at future asymtopia, the
{\it rhs} of eq. (\ref{dens}) would vanish, implying unbroken
supersymmetry despite a thermal spectrum of radiated particles. However,
even in the explicit form (\ref{dens}), it is not obvious how this
happens. Using eq. (\ref{haw}) and standard properties of density matrices, one
can compute
the {\it rhs} of (\ref{dens}) and obtain,
\begin{eqnarray}
\langle~ \delta_S {\bar Q}_{\dot \alpha}~\rangle~&=&~\epsilon^{\beta}~
\sum_k k_{\beta {\dot \alpha}}~(e^{4\pi |k|/ \kappa}~-~1)^{-1}~[~ (|t^B_k|^2~ +~
|t^F_k|^2)~\nonumber \\
&+&~ e^{2\pi |k|/ \kappa}~(|t^B_k|^2~-~ |t^F_k|^2) -2 |t^B_k|^2~|t^F_k|^2~ ]~.
\label{res}
\end{eqnarray}
Thus, whether the {\it rhs} of (\ref{res}) vanishes or not, depends on the nature
of the transmission coefficients (grey body factors) $|t^{B,F}_k|^2$. The
calculation of these coefficients depends on explicit solutions of the matter
field equations in specific
black hole backgrounds - a task we do not attempt here. However, we note that in the
limit $|t^B_k|^2 = |t^F_k|^2 = 1$, i.e., in the limit that the
potential barrier of the black hole for outgoing particles is {\it strictly}
reflectionless, $\langle~\delta_S {\bar Q}~\rangle~=~0$. Alternatively, the {\it
lhs} of eq. (\ref{res}) vanishes whenever,
\begin{equation}
|t^F_k|^2~~=~~{|t^B_k|^2~(e^{2\pi |k|/ \kappa} + 1) \over {(e^{2\pi |k| / \kappa}
-1) +2|t^B_k|^2 }}~. \label{nul}
\end{equation}
Of course, both the above requirements are non-generic: the first, namely that the
potential barrier is reflectionless, is extremely unrealistic. As for the second,
namely eq. (\ref{nul}), this is something which only detailed calculation of grey
body factors $|t^{B,F}_k|^2$ for specific black hole metrics can verify.
Observe, using eq. (\ref{haw}) that eq. (\ref{nul}) can be rewritten as
\begin{equation}
\langle~N_k^F~\rangle~~=~~{\langle~N_k^B~\rangle \over
(1~+~2\langle~N_k^B~\rangle)} ~. \label{ssrel}
\end{equation}
\section{Supersymmetry in Minkowski space at Finite Temperature}
A comparison, with the behaviour of the system of massless supersymmetric
bosons and fermions in flat spacetime at a finite temperature
$\beta^{-1}$, is in order. In this case, we have full
knowledge of the density operator as a function of the basic field
operators \cite{fey},
\begin{equation}
\rho~~=~~e^{-\beta~H}~/Tr~e^{-\beta H}~~, \label{temp}
\end{equation}
where, the Hamiltonian $H~=~H^B~(A^B,B^B)~+~H^F(A^F,B^F)$. Using eq. (\ref{trc}),
and recalling that there is no {\it explicit} supersymmetry breaking at ${\cal
I}^+$ so that $[{\bar Q}~,~e^{-\beta H}] =0$, we obtain,
\begin{equation}
\langle~ \delta_S {\bar Q}_{\dot \alpha}~\rangle~~=0~~.
\end{equation}
More explicitly, taking recourse once again to (\ref{enef}), one can actually
calculate the required vev: it is straightforward to see that the thermal average
\begin{equation}
\langle~\delta_S {\bar
Q}~\rangle_{\beta}~=~Tr~\{~\epsilon~H~e^{-\beta~H} \}
/Tr~e^{-\beta~H}~~. \label{thav} \end{equation}
Now,
\begin{equation}
Tr~\{~\epsilon~H~e^{-\beta~H} \}~=~-{d \over d
\beta}~Tr~\epsilon~e^{-\beta~H}~~. \label{der} \end{equation}
Therefore, to obtain preservation of supersymmetry at a non-zero
temperature, all we have to prove is that $Tr~\epsilon~e^{-\beta~H}$ is
independent of $\beta$.
To see this, we use the fact that the Hamiltonian $H$ is actually a sum of
an infinite number of bosonic and fermionic (spin 1/2) harmonic
oscillator Hamiltonians, each at a frequency $\omega_k~=~|k|$. Thus,
\begin{eqnarray}
Tr~\epsilon~e^{-\beta~H}~&=&~~\sum_k~\sum_{n^B_k~,~n^F_k}~\langle
n^B_k,n^F_k|~ \epsilon~e^{-\beta~H}~|n^B_k,n^F_k \rangle~ \nonumber \\
&=&~~\epsilon \sum_{k;n^B_k, n^F_k}~(-)^{n^F_k}~\exp
\{-\beta~\omega_k~(~n^B_k~+~n^F_k~)~\}~~\nonumber \\
&=&~~\sum_k~\sum_{n^B_k=0}^{\infty}~
\left(~e^{-\beta~n^B_k~\omega_k}~-~e^{-\beta~(n^B_k~+~1)~\omega_k}~\right)~~.
\label{qed}
\end{eqnarray}
It is clear that there is a term by term cancellation for each value of
$n^B_k$ in the {\it rhs} of eq. (\ref{qed}), {\it except} for the first
term for $n^B_k=0$, which of course is {independent of $\beta$}. Thus, {\it just
because bosons and fermions obey different statistics at a finite
temperature, it is hasty to conclude that supersymmetry is broken.} This
fact was first pointed out by L. van Hove \cite{vh} and subsequently by
other workers \cite{par}. Our formulation here makes only
implicit use of the Klein operator $(-)^{\hat {N^F}}$ in contrast to its
explicit use in those earlier papers.
\section{Conclusions}
The conditions which lead to $\langle~\delta_S {\bar Q}~\rangle~=~0$, namely that
either the potential barrier of the black hole is reflectionless, {\it or} eq.
(\ref{nul}) above holds, are both non-generic,
requiring calculation of grey body factors for specific black holes. Unlike in
section II where, by
regarding the supersymmetry parameter as a $c-number$ and blithely moving it
outside of vevs, we were able to show generically that supersymmetry is broken,
now it seems that the situation is actually more complicated. The comparison
with the flat space finite temperature case in section IV underlines this
feature fairly well: if none of the conditions for $\langle~\delta_S {\bar
Q}~\rangle~=~0$ hold, then the results do not guarantee that a black body
spectrum is all there is to black hole radiation. The disparity with the
flat space behaviour needs to be probed more thoroughly. We hope to report
on this in the near future.
It is important to point out that even if one suceeded in
establishing $\langle~\delta_S {\bar Q}~\rangle~=~0$, in principle there
could be other fermionic operators whose supersymmetry variations have
nonvanishing vevs. However, the simplicity of the model under
consideration makes it very unlikely that the foregoing will be challenged
in any manner. It is of course quite another question if interacting
fields are considered. The calculation of the vev will then have to done
perturbatively, in general, as one would analyze the flat space $T \neq 0$
situation. However, the latter situation has already been considered in
\cite{par}: appropriate use of the properties of the supersymmetry
parameter seems to preserve all supersymmetry Ward identities. So our
result should go through in that case as well.
\vglue .3in
I thank R. Kaul and A. Dasgupta for useful discussions.
|
\section*{}\list
{$^{\arabic{enumi}}$}{\settowidth\labelwidth{#1}\leftmargin\labelwidth
\advance\leftmargin\labelsep
\usecounter{enumi}}
\def\hskip .11em plus .33em minus .07em{\hskip .11em plus .33em minus .07em}
\sloppy\clubpenalty4000\widowpenalty4000
\sfcode`\.=1000\relax}
\let\Large=\large
\def\op#1{\mathop{\fam0 #1}\limits}
\newcommand{{\rm Id\,}}{{\rm Id\,}}
\newcommand{{\rm pr}}{{\rm pr}}
\newcommand{{\rm Ker\,}}{{\rm Ker\,}}
\newcommand{\nm}[1]{\mid {#1}\mid}
\newcommand{\dv}[1]{\op {#1}^\circ}
\newcommand{\begin{equation}}{\begin{equation}}
\newcommand{\end{equation}}{\end{equation}}
\newcommand{\begin{eqnarray}}{\begin{eqnarray}}
\newcommand{\end{eqnarray}}{\end{eqnarray}}
\newcommand{\begin{eqnarray*}}{\begin{eqnarray*}}
\newcommand{\end{eqnarray*}}{\end{eqnarray*}}
\newcommand{\begin{eqalph}}{\begin{eqalph}}
\newcommand{\end{eqalph}}{\end{eqalph}}
\newcommand{{\cal A}}{{\cal A}}
\newcommand{{\cal T}}{{\cal T}}
\newcommand{{\cal P}}{{\cal P}}
\newcommand{{\cal R}}{{\cal R}}
\newcommand{{\cal L}}{{\cal L}}
\newcommand{{\cal V}}{{\cal V}}
\newcommand{{\cal E}}{{\cal E}}
\newcommand{{\cal H}}{{\cal H}}
\newcommand{{\cal F}}{{\cal F}}
\newcommand{{\cal S}}{{\cal S}}
\newcommand{{\cal B}}{{\cal B}}
\newcommand{{\cal J}}{{\cal J}}
\newcommand{{\bf L}}{{\bf L}}
\newcommand{{\bf R}}{{\bf R}}
\newcommand{{\bf C}}{{\bf C}}
\newcommand{{\bf Z}}{{\bf Z}}
\newcommand{\alpha}{\alpha}
\newcommand{\beta}{\beta}
\newcommand{\delta}{\delta}
\newcommand{\lambda}{\lambda}
\newcommand{\Lambda}{\Lambda}
\newcommand{\phi}{\phi}
\newcommand{\omega}{\omega}
\newcommand{\Omega}{\Omega}
\newcommand{\mu}{\mu}
\newcommand{\nu}{\nu}
\newcommand{\gamma}{\gamma}
\newcommand{\Gamma}{\Gamma}
\newcommand{\epsilon}{\epsilon}
\newcommand{\varepsilon}{\varepsilon}
\newcommand{\theta}{\theta}
\newcommand{\Theta}{\Theta}
\newcommand{\vartheta}{\vartheta}
\newcommand{\sigma}{\sigma}
\newcommand{\wedge}{\wedge}
\newcommand{\widetilde}{\widetilde}
\newcommand{\widehat}{\widehat}
\newcommand{\overline}{\overline}
\newcommand{\partial}{\partial}
\newcommand{\op\longrightarrow}{\op\longrightarrow}
\newcommand{\longleftrightarrow}{\longleftrightarrow}
\newcommand{\otimes}{\otimes}
\newcommand{\approx}{\approx}
\newcommand{\nw}[1]{[{#1}]}
\newcommand{\rm Der}{\rm Der}
\renewcommand{\arabic{equationa}\alph{equation}}{\arabic{equation}}
\let\ssection=\section
\renewcommand{\section}{\setcounter{equation}{0}\ssection}
\newcounter{eqalph}
\newcounter{equationa}
\newcounter{theorem}
\newcounter{proposition}
\newcounter{lemma}
\newcounter{corollary}
\newcounter{definition}
\newcounter{remark}
\setcounter{theorem}{0}
\setcounter{remark}{0}
\setcounter{proposition}{0}
\setcounter{lemma}{0}
\setcounter{corollary}{0}
\setcounter{definition}{0}
\def\arabic{equationa}\alph{equation}{\arabic{equation}}
\newenvironment{eqalph}{\stepcounter{equation}
\setcounter{equationa}{\value{equation}}
\setcounter{equation}{0}
\def\arabic{equationa}\alph{equation}{\arabic{equationa}\alph{equation}}
\begin{eqnarray}}{\end{eqnarray}\setcounter{equation}{\value{equationa}}}
\def\arabic{definition}{\arabic{definition}}
\def\arabic{definition}{\arabic{definition}}
\def\arabic{definition}{\arabic{definition}}
\def\arabic{definition}{\arabic{definition}}
\def\arabic{remark}{\arabic{remark}}
\newenvironment{proof}{\noindent
{\it Proof:}}{\medskip}
\newenvironment{rem}{\refstepcounter{remark}\medskip\noindent{\it
Remark \arabic{remark}:}}{\medskip}
\newenvironment{theo}{\refstepcounter{definition}
\medskip\noindent{\it Theorem \thedefinition:}}{\medskip}
\newenvironment{prop}{\refstepcounter{definition}
\medskip\noindent{\it Proposition \thedefinition:}}{\medskip}
\newenvironment{lem}{\refstepcounter{definition}
\medskip\noindent{\it Lemma \thedefinition:}}{\medskip}
\newenvironment{cor}{\refstepcounter{definition}
\medskip\noindent{\it Corollary \thedefinition:}}{\medskip}
\hyphenation{ma-ni-fold La-gran-gi-ans di-men-si-o-nal
-di-men-si-o-nal La-gran-gi-an Ha-mil-to-ni-an multi-symplec-tic}
\begin{document}
\hbox{}
{\parindent=0pt
{\large\bf Covariant Hamiltonian field theory}
\bigskip
{\sc G.Giachetta and L.Mangiarotti}\footnote{Electronic mail:
<EMAIL>}
{\sl Department of Mathematics and Physics, University of Camerino, 62032
Camerino (MC), Italy}
\medskip
{\sc G. Sardanashvily}\footnote{Electronic mail:
<EMAIL>}
{\sl Department of Theoretical Physics,
Moscow State University, 117234 Moscow, Russia}
\bigskip
We study the relations between the equations of first order Lagrangian field
theory on fiber bundles and the covariant Hamilton equations on the
finite-dimensional polysymplectic phase space of covariant Hamiltonian
field theory. The main peculiarity of these Hamilton equations lies in the fact
that, for degenerate systems, they contain additional gauge fixing conditions.
We develop the BRST extension of the covariant
Hamiltonian formalism, characterized by a Lie superalgebra of
BRST and anti-BRST symmetries.
\bigskip
PACS numbers: 11.10.Ef, 11.30.Pb }
\bigskip
\noindent
{\bf I. INTRODUCTION}
\bigskip
As is well known, when applied to field theory, the familiar
symplectic techniques of mechanics take the form of instantaneous Hamiltonian
formalism on an infinite-dimensional phase space.
The finite-dimensional covariant Hamiltonian approach to field theory is
vigorously developed from the seventies in its multisymplectic and
polysymplectic variants.$^{1-3}$ Its final purpose
is the covariant Hamiltonian quantization of field theory.
In the framework of this approach, one deals with the
following types of PDEs: Euler--Lagrange and
Cartan equations in the Lagrangian formalism, Hamilton--De Donder
equations in multisymplectic Hamiltonian formalism, covariant Hamilton
equations and restricted Hamilton equations in polysymplectic
Hamiltonian formalism. If a Lagrangian is
hyperregular, all these PDEs are equivalent.
The present work
addresses degenerate semiregular and almost
regular Lagrangians. From the mathematical viewpoint, these notions of
degeneracy are particularly appropriate in order to study the relations between
the above-mentioned PDEs. From the physical one, Lagrangians of almost all
field theories are of these types.
To formulate our results, let us recall briefly some notions.
Given a fiber bundle
$Y\to X$, coordinated by $(x^\lambda,y^i)$, a first order Lagrangian $L$ is
defined as a horizontal density
\begin{equation}
L={\cal L}\omega: J^1Y\to\op\wedge^nT^*X, \quad \omega=dx^1\wedge\cdots dx^n, \quad n=\dim X,
\label{cmp1}
\end{equation}
on the affine jet bundle $J^1Y\to Y$, provided with the adapted coordinates
$(x^\lambda,y^i,y^i_\lambda)$. $J^1Y$ can be seen as a finite-dimensional
configuration space of fields represented by sections of $Y\to X$.
Given a Lagrangian
$L$ (\ref{cmp1}), the associated Euler--Lagrange equations
define both the equations of the variational
problem on $Y$ for $L$ and the kernel of the
Euler--Lagrange operator, which can be also introduced in an intrinsic
way as a coboundary element of the variational cochain complex.
The Cartan equations characterize
the variational problem on $J^1Y$ for the
Poincar\'e--Cartan form $H_L$, which is a horizontal Lepagean equivalent of
$L$ on $J^1Y\to Y$, i.e. $L=h_0(H_L)$, where $h_0$ is the horizontal
projection (\ref{cmp100}). At the same time, the Cartan equations can be seen
both as the kernel of the Euler--Lagrange--Cartan operator and the Hamilton
equations of the Lagrangian polysymplectic structure on
$J^1Y$. The Cartan equations are the Lagrangian
counterpart of covariant Hamilton equations.
The Hamilton-De Donder equations and the covariant Hamilton
equations are related to two different Legendre morphisms in the first
order calculus of variations.
Firstly, every
Poincar\'e--Cartan form
$H_L$ yields the Legendre morphism
$\widehat H_L$ of
$J^1Y$ to the homogeneous Legendre bundle
\begin{equation}
Z_Y=J^{1\star}Y = T^*Y\wedge(\op\wedge^{n-1}T^*X) \label{N41}
\end{equation}
which is the affine $\op\wedge^{n+1}$-valued dual of $J^1Y\to Y$,$^{1,4}$
and is treated as
a homogeneous finite-dimensional phase space of fields.
$Z_Y$ is provided with the canonical
exterior $n$-form
$\Xi_Y$ (\ref{N43}) and the multisymplectic form $d\Xi_Y$. If
$\widehat H_L(J^1Y)$ is an imbedded subbundle of $Z_Y\to Y$, the
pull-back of $\Xi_Y$ yields the Hamilton--De Donder
equations on $\widehat H_L(J^1Y)$.
If a Lagrangian
$L$ is almost regular, these equations are quasi-equivalent to the
Cartan equations, i.e., there is a surjection of the set of solutions
of the Cartan equations onto that of the Hamilton-De Donder equations.$^1$
Secondly, every Lagrangian $L$ defines the Legendre map $\widehat L$ of
$J^1Y$ to the Legendre bundle
\begin{equation}
\Pi=\op\wedge^nT^*X\op\otimes_YV^*Y\op\otimes_YTX, \label{00}
\end{equation}
provided with the holonomic coordinates $(x^\lambda,y^i,p^\lambda_i)$, which can be
seen as a finite-dimensional momentum phase space of fields.$^{5-9}$
The relationship between multisymplectic and polysymplectic phase spaces is
given by the exact sequence
\begin{equation}
0\op\longrightarrow\Pi\op\times_X\op\wedge^nT^*X\hookrightarrow Z_Y\op\longrightarrow\Pi\op\longrightarrow 0, \label{b417}
\end{equation}
where
\begin{equation}
\pi_{Z\Pi}:Z_Y\to \Pi\label{b418'}
\end{equation}
is a 1-dimensional affine bundle. Given any section
$h$ of $Z_Y\to\Pi$, the pull-back
\begin{equation}
H=h^*\Xi_Y= p^\lambda_i dy^i\wedge \omega_\lambda -{\cal H}\omega \label{b418}
\end{equation}
is a
polysymplectic Hamiltonian form on $\Pi$.$^{2-4}$
The Legendre $\Pi$ is equipped with the
canonical polysymplectic form $\Omega_Y$ (\ref{406}).$^{2,3}$
This form differs from those in
Refs. [4-6], and is globally defined. With the polysymplectic
form
$\Omega_Y$, one introduces Hamiltonian connections and covariant Hamilton
equations
\begin{eqalph}
&& y^i_\lambda=\partial^i_\lambda{\cal H}, \label{b4100a}\\
&& p^\lambda_{\lambda i}=-\partial_i{\cal H},\label{b4100b}
\end{eqalph}
which are PDEs on the phase space $\Pi$ defined by the kernel of the Hamilton
operator ${\cal E}_H$ (\ref{3.9}). If
$X={\bf R}$, covariant Hamiltonian formalism provides the adequate
Hamiltonian formulation of time-dependent mechanics.$^{10,11}$
In this case, $Z_Y=T^*Y$ and $\Pi=V^*Y$ are the
homogeneous and momentum phase spaces
of time-dependent mechanics, respectively.
It should be emphasized that
a Hamiltonian form $H$ (\ref{b418}) is the Poincar\'e--Cartan form of the
Lagrangian
\begin{equation}
L_H=h_0(H) = (p^\lambda_iy^i_\lambda - {\cal H})\omega \label{Q3}
\end{equation}
on the jet manifold $J^1\Pi$. It is the
the Poincar\'e--Cartan form (\ref{303})
of this Lagrangian. It follows that the Euler--Lagrange operator (\ref{305})
for $L_H$ is precisely the Hamilton operator ${\cal E}_H$ (\ref{3.9}) for
$H$ and, consequently, the Euler--Lagrange equations for
$L_H$ are equivalent to the Hamilton equations for $H$.
The Lagrangian $L_H$ plays a prominent role
in the path integral approach to quantization of Hamiltonian
systems.$^{12-14}$
The results of this paper demonstrate that polysymplectic Hamiltonian formalism
is not equivalent to the Lagrangian one, but can provide the adequate
description of degenerate field systems which do not necessarily possess
gauge symmetries.
We show that, if $r:X\to\Pi$ is a solution of the Hamilton equations for a
Hamiltonian form $H$ associated with a semiregular Lagrangian $L$ and if $r$
lives in the Lagrangian constraint space $\widehat L(J^1Y)$, then the
projection of $r$ onto $Y$ is a solution of the Euler--Lagrange
equations for $L$. The converse assertion is more intricate.
One needs a complete set of associated Hamiltonian forms
in order to exhaust all solutions of
the Euler--Lagrange equations for a degenerate Lagrangian.
It follows that the
covariant Hamilton equations contain additional conditions
in comparison with the Euler--Lagrange ones. In the case of
almost regular Lagrangians, one can introduce the constrained
Hamilton equations. They are weaker than the Hamilton equations
restricted to the Lagrangian constraint space, are
equivalent to the Hamilton--De Donder equations and, consequently, are
quasi-equivalent to the Cartan equations.
The detailed analysis
of degenerate quadratic Lagrangian systems in Section VI is appropriate
for application to many physical models. We
find a complete set of associated Hamiltonian forms. The
key point is the splitting of the configuration space $J^1Y$ into the dynamic
sector and the gauge one coinciding with the kernel of the Legendre map
$\widehat L$. As an immediate consequence of this splitting, one can separate a part
of the Hamilton equations independent of
momenta which play the role of gauge-type conditions, while other
equations restricted to the Lagrangian constraint space coincide with the
constrained Hamilton equations, and are quasi-equivalent to the Cartan
equations.
Thus, we observe that the main features in gauge
theory are not directly related to the gauge invariance condition, but are
common in all field models with degenerate quadratic Lagrangians. The
important peculiarity of the Hamiltonian description of these models lies in
the fact that, in comparison with a Lagrangian, an associated Hamiltonian
form $H$ and the Lagrangian $L_H$ (\ref{Q3}) contain gauge fixing terms.
Therefore we will construct the BRST extension of the Hamiltonian form $H$
(\ref{b418}) and the Lagrangian $L_H$ (\ref{Q3}) in order to provide them with
symmetries which lead, e.g., to the corresponding Slavnov identities under
quantization. This is a preliminary step towards the covariant Hamiltonian
quantization of degenerate systems.
A natural idea of the covariant Hamilton quantizations is also to generalize
the Poisson bracket in symplectic mechanics to multisymplectic or
polysymplectic manifolds and then to quantize it.$^{15}$ The main difficulty is
that the bracket must be globally defined. Let us note that
multisymplectic manifolds such as $(Z_Y,d\Xi_Y)$,
look rather promising for algebraic constructions since multisymplectic forms
are exterior forms.$^{16-18}$
Nevertheless, the above mentioned $X={\bf R}$ reduction of the covariant
Hamiltonian formalism leads to time-dependent mechanics, but not conservative
symplectic mechanics.
In this case, the momentum phase space
$V^*Y$, coordinated by $(t,y^i,p_i=\dot y_i)$, is endowed with the canonical
degenerate Poisson structure given by the bracket
\begin{equation}
\{f,g\}_V=\partial^if\partial_i g-\partial^ig\partial_i f, \qquad f,g\in C^\infty(V^*Y).
\label{c3}
\end{equation}
However, the Poisson bracket $\{{\cal H},f\}_V$ of a Hamiltonian ${\cal H}$ and
functions $f$ on the momentum phase space $V^*Y$ fails to be a well-behaved
entity because ${\cal H}$ is not a scalar
with respect to time-dependent transformations. In particular, the equality
$\{{\cal H},f\}_V=0$ is not preserved under such
transformations.$^{10,11}$ As a consequence, the evolution equation in
time-dependent mechanics is not reduced to a Poisson bracket. At the same
time, the Poisson bracket (\ref{c3}) leads to the following current algebra
bracket. Let
$u=u^i\partial_i$ be a vertical vector field on $Y\to{\bf R}$, and $J_u=u^ip_i$ the
corresponding symmetry current on $V^*Y$ along $u$. The symmetry
currents
$J_u$ constitute a Lie algebra with respect to the bracket
\begin{eqnarray*}
[J_u,J_{u'}]=\{J_u,J_{u'}\}_V= J_{[u,u']}.
\end{eqnarray*}
This current algebra bracket can be extended to the general polysymplectic case
as follows.
There is the canonical isomorphism
\begin{eqnarray*}
\overline\theta= p^\lambda_i\overline dy^i\wedge\omega_\lambda: \Pi\to
V^*Y\op\wedge_Y(\op\wedge^{n-1}T^*X).
\end{eqnarray*}
Let $u=u^i\partial_i$ be a vertical vector field on
$Y\to X$. The corresponding symmetry current (\ref{cmp129})
is a horizontal exterior $(n-1)$-form
\begin{equation}
J_u=u\rfloor\overline\theta=u^ip_i^\lambda\omega_\lambda \label{c1}
\end{equation}
on the Legendre bundle $\Pi$ (\ref{00}). The symmetry currents (\ref{c1})
constitute a Lie algebra with respect to the bracket
\begin{equation}
[J_u,J_{u'}]\op=^{\rm def} J_{[u,u']}. \label{c2}
\end{equation}
If $Y\to X$ is a vector bundle and $X$ is provided with a non-degenerate
metric $g$, the bracket (\ref{c2}) can be extended to any horizontal
$(n-1)$-forms $\phi=\phi^\alpha\omega_\alpha$ on $\Pi$ by the law
\begin{eqnarray*}
[\phi,\sigma]=g_{\alpha\beta}g^{\mu\nu}(\partial^i_\mu\phi^\alpha\partial_i\sigma^\beta-
\partial^i_\mu\sigma^\beta\partial_i\phi^\alpha)\omega_\nu.
\end{eqnarray*}
Similarly, the bracket of horizontal 1-forms on $\Pi$ is defined.$^{11}$
The bracket (\ref{c2}) looks promising for the current algebra quantization
of the covariant Hamiltonian formalism. We will use this bracket in order to
construct the algebra of supercurrents in the BRST extended Hamiltonian
formalism.
Note that, since the above mentioned Poisson bracket $\{{\cal H},f\}_V$ is not
preserved under time-dependent transformations, the standard BRST technique,
based on the Lie algebra of constraints, can not be applied in a
straightforward manner to time-dependent mechanics and covariant Hamiltonian
field theory. We generalize the BRST mechanics of
E.Gozzi and M.Reuter$^{11,12,19}$ in the terms of simple graded manifolds.
\bigskip
\noindent
{\bf II. TECHNICAL PRELIMINARIES}
\bigskip
All maps throughout the paper are smooth, while manifolds are real,
finite-dimensional, Hausdorff, second-countable and connected. A base
manifold $X$ is oriented.
Given a fiber bundle $Y\to X$ coordinated by $(x^\lambda,y^i)$, the
$s$-order jet manifold $J^sY$ is endowed with the adapted
coordinates $(x^\lambda,y^i_\Lambda)$, $0\leq\mid\Lambda\mid\leq s$, where
$\Lambda$ is a symmetric multi-index $(\lambda_k...\lambda_1)$, $\nm\Lambda=k$.
The repeated jet manifold $J^1J^1Y$ is coordinated by
$(x^\lambda,y^i,y^i_\lambda,\widehat y^i_\lambda,y^i_{\lambda\mu})$.
There are the canonical morphisms
\begin{eqnarray}
&&\lambda=dx^\lambda \otimes (\partial_\lambda + y^i_\lambda \partial_i):J^1Y\op\hookrightarrow_Y T^*X
\op\otimes_Y TY,\label{18}\\
&&S_1= (\widehat y_\lambda^i -y_\lambda^i)dx^\lambda\otimes\partial_i:J^1J^1Y\op\to_Y T^*X\op\otimes_Y
VY.\label{cmp18}
\end{eqnarray}
Exterior forms $\phi$ on a manifold $J^sY$, $s=0,1,\ldots$, are naturally
identified with their pull-backs onto $J^{s+1}Y$. There is the exterior algebra
homomorphism, called the horizontal projection,
\begin{equation}
h_0:\phi_\lambda dx^\lambda+ \phi_i^\Lambda dy^i_\Lambda \mapsto \phi_\lambda
dx^\lambda+\phi_i^\Lambda y^i_{\lambda+\Lambda}dx^\lambda \label{cmp100}
\end{equation}
which sends exterior forms on $J^sY$ onto the horizontal forms on
$J^{s+1}Y\to X$,
and vanishes on the contact forms $\theta^i_\Lambda=dy^i_\Lambda
-y^i_{\lambda+\Lambda}dx^\lambda$.
Note that the horizontal projection $h_0$ and the
pull-back operation with respect to bundle morphisms over $X$
mutually commute. Recall also the operators of
the total derivative
\begin{eqnarray*}
d_\lambda =
\partial_\lambda +y^i_{\lambda+\Lambda}\partial^\Lambda_i=\partial_\lambda +y^i_\lambda\partial_i + y^i_{\lambda\mu}\partial_i^\mu
+\cdots,
\end{eqnarray*}
and the horizontal differential
$d_H\phi=dx^\lambda\wedge d_\lambda\phi$ such that $h_0\circ d=d_H\circ h_0$.
We regard a connection on a fiber bundle $Y\to X$ as a
global section
\begin{equation}
\Gamma=dx^\lambda\otimes(\partial_\lambda +\Gamma^i_\lambda\partial_i) \label{cmp91}
\end{equation}
of the affine jet bundle $\pi^1_0:J^1Y\to Y$.$^{3,20}$
Sections of the underlying vector bundle $T^*X\op\otimes_YVY\to Y$ are called
soldering forms.
Every connection $\Gamma$ on a fiber bundle $Y\to X$
gives rise to the connection
\begin{equation}
V\Gamma=dx^\lambda\otimes(\partial_\lambda +\Gamma^i_\lambda\partial_i +\partial_j\Gamma^i_\lambda\dot y^j
\frac{\partial}{\partial\dot y^i}) \label{cmp43}
\end{equation}
on the fiber bundle $VY\to X$.
\bigskip
\noindent
{\bf III. LAGRANGIAN DYNAMICS}
\bigskip
We follow the first variational formula of the
calculus of variations.$^3$ Given a Lagrangian $L$ and its Lepagean
equivalent
$H_L$, this formula provides the canonical decomposition of the Lie derivative
of $L$ along a projectable vector field $u$ on $Y$ in accordance with the
variational problem. We restrict our consideration to
the Poincar\'e--Cartan form
\begin{equation}
H_L={\cal L}\omega +\pi^\lambda_i\theta^i\wedge\omega_\lambda, \quad \pi^\lambda_i=\partial^\lambda_i{\cal L}, \quad
\omega_\lambda=\partial_\lambda\rfloor\omega.
\label{303}
\end{equation}
In contrast with other Lepagean equivalents,
$H_L$ is a horizontal form on the affine jet bundle $J^1Y\to Y$. Moreover, it
is the Lagrangian counterpart of polysymplectic Hamiltonian forms (see the
relations (\ref{4.9}) and (\ref{2.32}) below). The first variational formula
reads
\begin{equation}
{\bf L}_{J^1u}L=
u_V\rfloor {\cal E}_L + d_Hh_0(u\rfloor H_L), \label{C30}
\end{equation}
where $u_V=(u\rfloor\theta^i)\partial_i$ and
\begin{equation}
{\cal E}_L=
(\partial_i- d_\lambda\partial^\lambda_i){\cal L} \theta^i\wedge\omega: J^2Y\to T^*Y\wedge(\op\wedge^nT^*X) \label{305}
\end{equation}
is the Euler--Lagrange operator associated with $L$. The kernel
of ${\cal E}_L$ defines the Euler--Lagrange equations on
$Y$ given by the coordinate relations
\begin{equation}
(\partial_i- d_\lambda\partial^\lambda_i){\cal L}=0. \label{b327}
\end{equation}
Solutions of these equations are
critical sections of the
variational problem for the Lagrangian $L$.
\begin{rem}
The first variational formula (\ref{C30}) also
provides the Lagrangian conservation laws.$^{3,21}$ On-shell, we have the
weak identity
\begin{eqnarray*}
{\bf L}_{J^1u}L\approx d_Hh_0(u\rfloor H_L),
\end{eqnarray*}
and, if ${\bf L}_{J^1u}L=0,$
the weak conservation law
\begin{equation}
0\approx d_Hh_0(u\rfloor H_L)= - d_\lambda {\cal T}^\lambda \omega \label{K4}
\end{equation}
of the symmetry current
\begin{equation}
{\cal T} =-h_0(u\rfloor H_L)={\cal T}^\lambda\omega_\lambda =-[\pi^\lambda_i(u^\mu y^i_\mu-u^i
)-u^\lambda{\cal L}]\omega_\lambda \label{Q30}
\end{equation}
along the vector field $u$.
\end{rem}
Instead of the variational problem on $Y$ for a Lagrangian $L$,
one can consider that on $J^1Y$ for the Poincar\'e--Cartan form
$H_L$ (\ref{303}).
Critical sections $\overline s: X\to J^1Y$ of this variational problem
satisfy the relation
\begin{equation}
\overline s^*(u\rfloor dH_L)=0 \label{C28}
\end{equation}
for all vertical vector fields $u$ on $J^1Y\to X$. This relation defines the
Cartan equations on $J^1Y$.
We regain these equations in another way.$^3$
Let us consider the above-mentioned Legendre map
\begin{eqnarray*}
\widehat L:J^1Y \op\to_Y \Pi, \qquad p^\lambda_i\circ\widehat L =\pi^\lambda_i.
\end{eqnarray*}
The Legendre bundle $\Pi$ (\ref{00}) is equipped with the canonical
tangent-valued Liouville form $\theta$ (\ref{2.4}). Its pull-back on $J^1Y$ by
$\widehat L$ is
\begin{eqnarray*}
\theta_L = \widehat L^*\theta = - \pi_i^\lambda dy^i\wedge\omega\otimes \partial_\lambda.
\end{eqnarray*}
We construct the
reduced Lagrangian
\begin{equation}
\overline L = L - S_1\rfloor\theta_L = ({\cal L} + (\widehat y_\lambda^i - y_\lambda^i)\pi_i^\lambda)\omega
\label{cmp80}
\end{equation}
on $J^1J^1Y$ (see the notation (\ref{cmp18})). The
associated Euler--Lagrange operator, called the Euler--Lagrange--Cartan
operator for $L$, reads
\begin{eqnarray}
&& {\cal E}_{\overline L} : J^1J^1Y\to T^*J^1Y\wedge(\op\wedge^n T^*X), \nonumber \\
&& {\cal E}_{\overline L} = [(\partial_i{\cal L} - \widehat d_\lambda\pi_i^\lambda
+ \partial_i\pi_j^\lambda(\widehat y_\lambda^j - y_\lambda^j))dy^i + \partial_i^\lambda\pi_j^\mu(\widehat
y_\mu^j - y_\mu^j) dy_\lambda^i]\wedge \omega, \label{2237} \\
&&\widehat d_\lambda=\partial_\lambda +\widehat y^i_\lambda\partial_i +y^i_{\lambda\mu}\partial_i^\mu. \nonumber
\end{eqnarray}
This is the Lagrangian counterpart of the polysymplectic Hamilton operator
(see the relation (\ref{b4.1000}) below). Its kernel
${\rm Ker\,}{\cal E}_{\overline L}\subset J^1J^1Y$ is given exactly by the Cartan equations
\begin{eqalph}
&& \partial_i^\lambda\pi_j^\mu(\widehat y_\mu^j - y_\mu^j)=0, \label{b336a}\\
&& \partial_i {\cal L} - \widehat d_\lambda\pi_i^\lambda
+ (\widehat y_\lambda^j - y_\lambda^j)\partial_i\pi_j^\lambda=0. \label{b336b}
\end{eqalph}
Since ${\cal E}_{\overline L}\mid_{J^2Y}={\cal E}_L$, the Cartan equations (\ref{b336a}) --
(\ref{b336b}) are equivalent to the Euler--Lagrange equations
(\ref{b327}) on integrable sections of
$J^1Y\to X$. These equations are equivalent
in the case of regular Lagrangians.
With the Poincar\'e--Cartan form $H_L$ (\ref{303}),
we have the Legendre morphism
\begin{eqnarray*}
\widehat H_L: J^1Y\op\to_Y Z_Y, \qquad
(p^\mu_i, p)\circ\widehat H_L =(\pi^\mu_i, {\cal L}-\pi^\mu_i y^i_\mu ),
\end{eqnarray*}
where the fiber bundle $Z_Y$ (\ref{N41}) is endowed with holonomic
coordinates $(x^\lambda,y^i,p^\lambda_i,p)$.
It is readily observed that
\begin{equation}
\widehat L=\pi_{Z\Pi}\circ \widehat H_L. \label{LZP}
\end{equation}
Owing to the monomorphism $Z_Y\hookrightarrow \op\wedge^nT^*Y$, the bundle
$Z_Y$ is equipped with the pull-back
\begin{equation}
\Xi_Y= p\omega + p^\lambda_i dy^i\wedge\omega_\lambda \label{N43}
\end{equation}
of the canonical form $\Theta$ on
$\op\wedge^nT^*Y$ whose exterior differential $d\Theta$ is the $n$-multisymplectic
form in the sense of Martin.$^{22}$
Let $Z_L=\widehat H_L(J^1Y)$ be an imbedded subbundle
$i_L:Z_L\hookrightarrow Z_Y$ of $Z_Y\to Y$.
It is provided with the pull-back
De Donder form $i^*_L\Xi_Y$. We have
\begin{equation}
H_L=\widehat H_L^*\Xi_L=\widehat H_L^*(i_L^*\Xi_Y). \label{cmp14}
\end{equation}
By analogy with the Cartan equations
(\ref{C28}), the Hamilton--De Donder equations for sections $\overline r$ of
$Z_L\to X$ are written as
\begin{equation}
\overline r^*(u\rfloor d\Xi_L)=0 \label{N46}
\end{equation}
where $u$ is an arbitrary vertical vector field on
$Z_L\to X$. To obtain an explicit form of these equations, one should
substitute solutions
$(y^i_\lambda$, ${\cal L})$
of the equations
\begin{equation}
p^\lambda_i=\pi^\lambda_i, \qquad
p={\cal L} -\pi^\lambda_i y^i_\lambda \label{N60}
\end{equation}
in the Cartan equations.
However, if a Lagrangian $L$ is
degenerate, the equations (\ref{N60}) may admit different solutions
or no solution at all. Something more is said in the following theorem.$^1$
\begin{theo}\label{ddd} Let the Legendre morphism
$\widehat H_L:J^1Y\to Z_L$ be a submersion. Then a section $\overline s$ of $J^1Y\to X$
is a solution of the Cartan equations (\ref{C28}) if and only if $\widehat
H_L\circ\overline s$ is a solution of the Hamilton--De Donder equations
(\ref{N46}), i.e., Cartan and Hamilton--De Donder equations are
quasi-equivalent.
\end{theo}
\bigskip
\noindent
{\bf IV. COVARIANT HAMILTONIAN DYNAMICS}
\bigskip
Given a fiber bundle $Y\to X$, let $\Pi$ be the
Legendre bundle (\ref{00}). Holonomic coordinates
$(x^\lambda ,y^i,p^\lambda_i)$ on $\Pi$ are compatible with its composite fibration
\begin{eqnarray*}
\pi_{\Pi X}=\pi\circ\pi_{\Pi Y}:\Pi\to Y\to X.
\end{eqnarray*}
We have the canonical bundle monomorphism
\begin{equation}
\theta =-p^\lambda_idy^i\wedge\omega\otimes\partial_\lambda
:\Pi\op\hookrightarrow_Y\op\wedge^{n+1}T^*Y\op\otimes_Y TX, \label{2.4}
\end{equation}
called the tangent-valued Liouville form on $\Pi$.
It should be emphasized that the exterior differential $d$ can not be applied
to the tangent-valued form (\ref{2.4}). At the same time, there is a unique
$TX$-valued $(n+2)$-form
$\Omega_Y$ on
$\Pi$ such that the relation
\begin{eqnarray*}
\Omega_Y\rfloor\phi =-d(\theta\rfloor\phi)
\end{eqnarray*}
holds for any exterior 1-form $\phi$ on $X$.$^3$
This form, called the polysymplectic form, is given by the coordinate
expression
\begin{equation}
\Omega_Y =dp_i^\lambda\wedge dy^i\wedge \omega\otimes\partial_\lambda. \label{406}
\end{equation}
As was mentioned above, every section $h$ of the fiber bundle (\ref{b418'})
defines the pull-back (\ref{b418}) of the canonical form $\Xi_Y$ (\ref{N43}),
called a Hamiltonian form on the Legendre bundle $\Pi$.
\begin{prop}\label{sphamf}
Hamiltonian forms on $\Pi$ constitute a non-empty affine space modelled
over the linear space of horizontal densities $\widetilde H=\widetilde{{\cal H}}\omega$
on $\Pi\to X$.
\end{prop}
\begin{proof}
This is an immediate consequence of the fact that
(\ref{b418'}) is an affine bundle modelled over the pull-back vector bundle
$\Pi\op\times_X\op\wedge^nT^*X\to \Pi$.
\end{proof}
\begin{lem}$^{2,4}$ \label{cmp2} Every connection
$\Gamma$ (\ref{cmp91}) on $Y\to X$ yields the splitting
\begin{eqnarray*}
h_\Gamma: \overline dy^i \mapsto dy^i- \Gamma^i_\lambda dx^\lambda
\end{eqnarray*}
of the exact sequence (\ref{b417}) and, as a consequence, defines the
Hamiltonian form
\begin{eqnarray}
H_\Gamma =h_\Gamma^*\Xi_Y =p^\lambda_i dy^i\wedge\omega_\lambda -p^\lambda_i\Gamma^i_\lambda\omega. \label{3.6}
\end{eqnarray}
\end{lem}
Proposition \ref{sphamf} and Lemma \ref{cmp2} lead to the
following
\begin{cor}\label{hams}
Given a connection $\Gamma$ on $Y\to X$, every
Hamiltonian form $H$ admits the decomposition
\begin{equation}
H=H_\Gamma -\widetilde H_\Gamma =p^\lambda_idy^i\wedge\omega_\lambda
-p^\lambda_i\Gamma^i_\lambda\omega-\widetilde{{\cal H}}_\Gamma\omega. \label{4.7}
\end{equation}
\end{cor}
\begin{rem}
The physical meaning of the splitting (\ref{4.7}) is illustrated by
the fact that, in the case of $X={\bf R}$, $\widetilde{\cal H}_\Gamma$ is exactly the energy of
a mechanical system with respect to the reference frame $\Gamma$.$^{10,11}$
\end{rem}
We will mean by a Hamiltonian map any bundle morphism
\begin{equation}
\Phi=dx^\lambda\otimes(\partial_\lambda +\Phi^i_\lambda\partial_i):\Pi\op\to_Y J^1Y.
\label{2.7}
\end{equation}
In particular, let $\Gamma$ be a connection on $Y\to X$. Then, the
composition
\begin{equation}
\widehat\Gamma=\Gamma\circ\pi_{\Pi Y}=dx^\lambda\otimes (\partial_\lambda +\Gamma^i_\lambda\partial_i):\Pi\to Y\to
J^1Y, \label{b420}
\end{equation}
is a Hamiltonian map. Conversely,
every Hamiltonian map $\Phi$ yields
the associated connection $\Gamma_\Phi =\Phi\circ\widehat 0$
on $Y\to X$, where $\widehat 0$ is the global zero section of the
Legendre bundle $\Pi\to Y$. In particular, we have
$\Gamma_{\widehat\Gamma}=\Gamma$. The following two facts will be used in the sequel.
\begin{prop}$^3$
Every Hamiltonian form $H$ (\ref{b418}) yields the Hamiltonian
map $\widehat H$ such that
\begin{equation}
y_\lambda^i\circ\widehat H=\partial^i_\lambda{\cal H}. \label{415}
\end{equation}
\end{prop}
\begin{prop} Every Hamiltonian map
(\ref{2.7}) defines the Hamiltonian form
\begin{eqnarray*}
H_\Phi=\Phi\rfloor\theta =p^\lambda_idy^i\wedge\omega_\lambda -p^\lambda_i\Phi^i_\lambda\omega.
\end{eqnarray*}
\end{prop}
\begin{proof}
Given an arbitrary connection $\Gamma$ on the fiber bundle $Y\to X$, the
corresponding Hamiltonian map (\ref{b420}) defines the form $\widehat\Gamma\rfloor\theta$
which is exactly the Hamiltonian form $H_\Gamma$ (\ref{3.6}). Since $\Phi-\widehat\Gamma$
is a $VY$-valued basic 1-form on $\Pi\to X$, $H_\Phi-H_\Gamma$ is a
horizontal density on $\Pi$. Then the result follows from Proposition
\ref{sphamf}. Note that $H=H_{\widehat H}$ iff $H=H_\Gamma$ (\ref{3.6}).
\end{proof}
Let $J^1\Pi$ be the first order jet
manifold of
$\Pi\to X$. It is equipped with the adapted coordinates
$( x^\lambda ,y^i,p^\lambda_i,y^i_\mu,p^\lambda_{\mu i})$ such that
$y^i_\mu\circ J^1\pi_{\Pi Y}=y^i_\mu$.
A connection
\begin{equation}
\gamma =dx^\lambda\otimes(\partial_\lambda +\gamma^i_\lambda\partial_i
+\gamma^\mu_{\lambda i}\partial^i_\mu) \label{cmp33}
\end{equation}
on $\Pi\to X$ is called a Hamiltonian
connection if the exterior form $\gamma\rfloor\Omega_Y$
is closed.
A Hamiltonian connection $\gamma$ is said to be associated with
a Hamiltonian form $H$ if it obeys the condition
\begin{eqnarray}
&& \gamma\rfloor\Omega_Y= dH, \label{cmp3}\\
&& \gamma^i_\lambda =\partial^i_\lambda{\cal H}, \qquad
\gamma^\lambda_{\lambda i}=-\partial_i{\cal H}. \label{3.10}
\end{eqnarray}
\begin{theo}\label{loch}
For every Hamiltonian connection $\gamma$, there exists a local Hamiltonian form
$H$ on a neighbourhood of any point $q\in\Pi$ such that the equation
(\ref{cmp3}) holds.
\end{theo}
\begin{proof} If $\gamma\rfloor\Omega_Y$ is closed, there is a
contractible neighbourhood
$U$ of a point $q\in\Pi$ which belongs to a holonomic coordinate chart
$(x^\lambda,y^i,p^\lambda_i)$ and where the local form $\gamma\rfloor\Omega_Y$ is exact.
We have
\begin{equation}
\gamma\rfloor\Omega_Y =dH=dp^\lambda_i\wedge dy^i\wedge\omega_\lambda -
(\gamma^i_\lambda dp^\lambda_i -\gamma^\lambda_{\lambda i}dy^i)\wedge\omega \label{cmp4}
\end{equation}
on $U$. It is readily observed that the second term in the right-hand side of
this equality is also an exact form on $U$. By virtue of the relative
Poincar\'e lemma, it can be brought into the form $d{\cal H}\wedge\omega$ where ${\cal H}$ is a
local function on $U$. Then the form $H$ in the expression
(\ref{cmp4}) reads
\begin{eqnarray*}
H=p^\lambda_idy^i\wedge\omega_\lambda -{\cal H}\omega.
\end{eqnarray*}
Using Corollary \ref{hams}, one can easily show that this is a
Hamiltonian form on $U$.
\end{proof}
\begin{theo}
Every Hamiltonian form has an associated
Hamiltonian connection.
\end{theo}
\begin{proof}
Given a Hamiltonian form $H$, let us consider the first
order differential operator
\begin{eqnarray}
&& {\cal E}_H :J^1\Pi\to T^*\Pi\wedge(\op\wedge^n T^*X),\nonumber \\
&& {\cal E}_H=dH-\lambda\rfloor\Omega_Y=[(y^i_\lambda-\partial^i_\lambda{\cal H}) dp^\lambda_i
-(p^\lambda_{\lambda i}+\partial_i{\cal H}) dy^i]\wedge\omega, \label{3.9}
\end{eqnarray}
on $\Pi$ where $\lambda$ is the canonical monomorphism (\ref{18}). It
is called the Hamilton operator associated with $H$. A glance at the expression
(\ref{3.9}) shows that this operator is an affine morphism over
$\Pi$ of constant rank. It follows that its kernel is an affine
closed imbedded subbundle of the jet bundle $J^1\Pi\to\Pi$. This subbundle
has a global section $\gamma$ which is a connection on $\Pi\to X$. This connection
obeys the equation (\ref{cmp3}).
\end{proof}
It should be emphasized that, if $n>1$, there is a set of
Hamiltonian connections associated with the same Hamiltonian form
$H$. They differ from each other in soldering forms $\sigma$ on $\Pi\to
X$ which obey the equation $\sigma\rfloor\Omega_Y=0$.
\begin{prop}
Every Hamiltonian connection $\gamma$ associated with a Hamiltonian form $H$
satisfies the relation
\begin{equation}
J^1\pi_{\Pi Y}\circ \gamma= \widehat H. \label{4.109}
\end{equation}
\end{prop}
\begin{proof}
The proof is based on the expressions (\ref{415}) and (\ref{3.10}).
\end{proof}
Being a closed subbundle of the jet
bundle
$J^1\Pi\to X$, the kernel of the Hamilton operator ${\cal E}_H$
(\ref{3.9}) defines first order Hamilton equations (\ref{b4100a}) --
(\ref{b4100b}) on the Legendre bundle
$\Pi$. Every integral section
$J^1r=\gamma\circ r$
of a Hamiltonian connection $\gamma$ associated with a Hamiltonian form $H$ is
obviously a solution of the Hamilton equations (\ref{b4100a}) --
(\ref{b4100b}). Conversely, a solution of the Hamilton equations
(\ref{b4100a}) -- (\ref{b4100b}) is
a section $r$ of $\Pi\to X$ such that
its jet prolongation $J^1r$ lives in ${\rm Ker\,}{\cal E}_H$.
If $r:X\to\Pi$
is a global solution, there exists an extension of the local section
$J^1r: r(X) \to J^1\Pi$
to a Hamiltonian connection which has $r$ as an integral
section. Substituting $J^1r$ in (\ref{4.109}), we obtain the equality
\begin{equation}
J^1(\pi_{\Pi Y}\circ r)= \widehat H\circ r, \label{N10}
\end{equation}
which is the coordinate-free form of the Hamilton equations (\ref{b4100a}).
Nevertheless, it may happen that the
Hamilton equations
have no solution through a given point $q\in \Pi$.
\begin{rem}
The Hamilton equations can be introduced
without appealing to the Hamilton operator.
As was for the Cartan equations (\ref{C28}),
they are equivalent to the condition
\begin{equation}
r^*(u\rfloor dH)= 0 \label{N7}
\end{equation}
for any vertical vector field $u$ on $\Pi\to X$.
\end{rem}
\bigskip
\noindent
{\bf V. LAGRANGIAN AND HAMILTONIAN DEGENERATE SYSTEMS}
\bigskip
Let us study the relations between
Hamilton and Euler--Lagrange equations when
a Lagrangian is degenerate.
Their main peculiarity lies in the fact that
there is a set of
Hamiltonian forms associated with the same degenerate Lagrangian.
\begin{rem} \label{cmp7}
Let us recall the coordinate expressions
\begin{eqnarray}
&& (y^i_\mu,\widehat y^i_\lambda,y^i_{\lambda\mu})\circ
J^1\widehat H=(\partial^i_\mu{\cal H},y^i_\lambda,d_\lambda\partial^i_\mu{\cal H}),\label{cmp5}\\
&&(p^\lambda_i, y^i_\mu,p^\lambda_{\mu i})\circ J^1\widehat L=
(\pi^\lambda_i,\widehat y^i_\mu, \widehat d_\mu\pi^\lambda_i). \label{cmp6}
\end{eqnarray}
In particular, if $\gamma$ is a Hamiltonian connection for a
Hamiltonian form $H$, we obtain from (\ref{N10}) and (\ref{cmp5}) that
the composition $J^1\widehat H\circ\gamma$ takes its
values
into the sesquiholonomic jet bundle $\widehat J^2Y$.
\end{rem}
A Hamiltonian form $H$ is said to be associated
with a Lagrangian $L$ if $H$ satisfies the relations
\begin{eqalph}
&&\widehat L\circ\widehat H\circ \widehat L=\widehat L,\label{2.30a} \\
&&H=H_{\widehat H}+\widehat H^*L. \label{2.30b}
\end{eqalph}
A glance at the relation (\ref{2.30a}) shows that $\widehat
L\circ\widehat H$ is the projector
\begin{equation}
p^\mu_i(q)=\partial^\mu_i{\cal L} (x^\mu,y^i,\partial^j_\lambda{\cal H}(q)), \qquad q\in N_L,
\label{b481'}
\end{equation}
from $\Pi$ onto the Lagrangian constraint space $N_L=\widehat L( J^1Y)$.
Accordingly, $\widehat H\circ\widehat L$ is the projector from $J^1Y$ onto $\widehat
H(N_L)$.
\begin{lem} \label{cmp110}
Any Hamiltonian form $H$ associated with a Lagrangian $L$ obeys
the relation
\begin{equation}
H\mid_{N_L}=\widehat H^*H_L\mid_{N_L}, \label{4.9}
\end{equation}
where $H_L$ is the Poincar\'e--Cartan form (\ref{303}).
\end{lem}
\begin{proof}
The relation (\ref{2.30b}) takes the coordinate form
\begin{equation}
{\cal H}=p^\mu_i\partial^i_\mu{\cal H}-{\cal L}(x^\mu,y^i,\partial^j_\lambda{\cal H}). \label{b481}
\end{equation}
Substituting (\ref{b481'}) and (\ref{b481}) in (\ref{b418}), we
obtain the relation (\ref{4.9}).
\end{proof}
Something more can be said in the case of semiregular Lagrangians.
A Lagrangian $L$ is called semiregular if the pre-image $\widehat L^{-1}(q)$ of
any point $q\in N_L$ is a connected submanifold of $J^1Y$. The following fact
will be used in the sequel.
\begin{lem} \label{3.22}
The
Poincar\'e--Cartan form $H_L$ for a
semiregular Lagrangian $L$ is constant on the connected
pre-image $\widehat L^{-1}(q)$ of any point $q\in N_L$.
\end{lem}
\begin{proof}
Let $u$ be a vertical vector field on the affine jet bundle $J^1Y\to
Y$ which takes its values into the kernel of the
tangent map $T\widehat L$ to $\widehat L$. Then ${\bf L}_u H_L=0$.
\end{proof}
An immediate consequence of this fact is the following assertion.
\begin{prop} \label{3.22'} All Hamiltonian forms associated
with a semiregular Lagrangian $L$ coincide with each
other on the Lagrangian constraint space $N_L$,
and the Poincar\'e--Cartan form $H_L$ (\ref{303})
for $L$ is the pull-back
\begin{eqnarray}
&& H_L=\widehat L^*H, \label{2.32}\\
&& (\pi^\lambda_iy^i_\lambda-{\cal L})\omega={\cal H}(x^\mu,y^j,\pi^\mu_j)\omega,\nonumber
\end{eqnarray}
of any such a Hamiltonian form $H$.
\end{prop}
\begin{proof}
Given a
vector $v\in T_q\Pi$, the value
$T\widehat H(v)\rfloor H_L(\widehat H(q))$ is the same for all Hamiltonian maps
$\widehat H$ satisfying the relation (\ref{2.30a}). Then the results follow from
the relation (\ref{4.9}).
\end{proof}
Proposition \ref{3.22'} enables us to connect Euler--Lagrange and Cartan
equations for a semiregular Lagrangian $L$ with the Hamilton
equations for Hamiltonian forms associated with $L$.
\begin{theo}\label{3.23}
Let a section $r$ of $\Pi\to X$
be a solution of the Hamilton equations (\ref{b4100a}) -- (\ref{b4100b})
for a Hamiltonian form $H$ associated with a semiregular
Lagrangian $L$. If $r$ lives in the constraint space $N_L$, the
section $s=\pi_{\Pi Y}\circ r$
of $Y\to X$ satisfies the Euler--Lagrange
equations (\ref{b327}), while $\overline s=\widehat H\circ r$ obeys the Cartan equations
(\ref{b336a}) -- (\ref{b336b}).
\end{theo}
\begin{proof}
Acting by the exterior differential on the relation
(\ref{2.32}), we obtain the relation
\begin{equation}
(y^i_\lambda-\partial^i_\lambda{\cal H}\circ\widehat L)d\pi^\lambda_i\wedge\omega-(\partial_i{\cal L}+ \partial_i({\cal H}\circ\widehat
L))dy^i\wedge\omega=0 \label{b4160}
\end{equation}
which is equivalent to the system of equalities
\begin{eqnarray*}
&&\partial^\lambda_i\pi^\mu_j(y^j_\mu -\partial^j_\mu{\cal H}\circ\widehat L)=0, \\
&& \partial_i\pi^\mu_j(y^j_\mu -\partial^j_\mu{\cal H}\circ\widehat L) -(\partial_i{\cal L} +(\partial_i{\cal H})\circ\widehat
L)=0.
\end{eqnarray*}
Using these equalities and the expression (\ref{cmp6}), one can easily see that
\begin{equation}
{\cal E}_{\overline L}=(J^1\widehat L)^*{\cal E}_H, \label{b4.1000}
\end{equation}
where ${\cal E}_{\overline L}$ is the Euler--Lagrange--Cartan operator (\ref{2237}).
Let $r$ be a section of $\Pi\to X$ which lives in the Lagrangian constraint
space $N_L$, and
$\overline s=\widehat H\circ r$. Then we have
\begin{eqnarray*}
r=\widehat L\circ \overline s, \qquad J^1r=J^1\widehat L\circ J^1\overline s.
\end{eqnarray*}
If $r$ is a solution of the Hamilton equations, the exterior form ${\cal E}_H$
vanishes on $J^1r(X)$. Hence, the pull-back form ${\cal E}_{\overline
L}=(J^1\widehat L)^*{\cal E}_H$ vanishes on $J^1\overline s(X)$. It follows that
$\overline s$ obeys the Cartan equations (\ref{b336a}) -- (\ref{b336b}). We obtain
from the equality (\ref{N10}) that $\overline s= J^1s$, $s=\pi_{\Pi Y}\circ r$.
Hence, $s$ is a solution of the Euler--Lagrange equations.
\end{proof}
The same result can be obtained from the relation
\begin{equation}
\overline L=(J^1\widehat L)^*L_H \label{cmp83}
\end{equation}
where $\overline L$ is the Lagrangian (\ref{cmp80}) on $J^1J^1Y$ and $L_H$ is the
Lagrangian (\ref{Q3}) on $J^1\Pi$.
\begin{theo}\label{3.24} Given a semiregular Lagrangian $L$,
let a section $\overline s$ of the jet bundle
$J^1Y\to X$ be a solution of the
Cartan equations (\ref{b336a}) -- (\ref{b336b}).
Let $H$ be a Hamiltonian form associated with $L$, and let $H$ satisfy the
relation
\begin{equation}
\widehat H\circ \widehat L\circ \overline s=J^1(\pi^1_0\circ\overline s).\label{2.36}
\end{equation}
Then, the section $r=\widehat L\circ \overline s$
of the Legendre bundle $\Pi\to X$ is a solution of the
Hamilton equations (\ref{b4100a}) -- (\ref{b4100b}) for $H$.
\end{theo}
\begin{proof} The Hamilton equations (\ref{b4100a}) hold by
virtue of the condition (\ref{2.36}). Substituting $\widehat L\circ \overline s$
in the Hamilton equations (\ref{b4100b}) and
using the relations
(\ref{b4160}) and (\ref{2.36}), we come to the Cartan equations
(\ref{b336b}) for $\overline s$ as follows:
\begin{eqnarray*}
&& \widehat d_\lambda \pi^\lambda_i\circ\overline s +(\partial_i{\cal H})\circ\widehat L\circ \overline s=
\widehat d_\lambda \pi^\lambda_i\circ\overline s + (\overline s^j_\mu
-\partial^j_\mu{\cal H}\circ\widehat L\circ \overline s)\partial_i\pi^\mu_j\circ\overline s -\partial_i{\cal L}\circ\overline s=
\\ && \qquad \widehat d_\lambda \pi^\lambda_i\circ\overline s- (\partial_\mu \overline s^j -\overline
s^j_\mu)\partial_i\pi^\mu_j\circ\overline s -\partial_i{\cal L}\circ\overline s=0.
\end{eqnarray*}
\end{proof}
\begin{rem} \label{cmp9}
Since $\widehat H\circ \widehat L$ in Theorem (\ref{3.24}) is a projection operator,
the condition (\ref{2.36}) implies that the solution $\overline s$ of the Cartan
equations is actually an integrable section $\overline s=J^1s$ where $s$ is a
solution of the Euler--Lagrange equations.
Theorems \ref{3.23} and \ref{3.24} show that, if a solution of the Cartan
equations provides a solution of the covariant Hamilton equations, it is
necessarily a solution of the Euler--Lagrange equations. In fact, the
relation (\ref{b4.1000}) gives more than it is needed for proving Theorem
\ref{3.23}. Using this relation, one can justify that, if $\gamma$ is a Hamiltonian
connection for a Hamiltonian form $H$ associated with a semiregular Lagrangian
$L$, then the composition $J^1\widehat H\circ\gamma\circ\widehat L$ takes its values in
${\rm Ker\,} {\cal E}_{\overline L}\cap\widehat{J^2Y}$ (see Remark \ref{cmp7}), i.e., this is a local
sesquiholonomic Lagrangian connection on
$\widehat H(N_L)$.$^3$ A converse of this assertion, however, fails to
be true in the case of semiregular Lagrangians.
Let a Lagrangian $L$ be
hyperregular, i.e., the Legendre map $\widehat L$
is a diffeomorphism. Then $\widehat L^{-1}$ is a Hamiltonian map, and there is a
unique Hamiltonian form
\begin{equation}
H=H_{\widehat L^{-1}}+\widehat L^{-1*}L \label{cmp8}
\end{equation}
associated with $L$. In this case, both the relation (\ref{b4.1000}) and the
converse one
\begin{eqnarray*}
{\cal E}_H=(J^1\widehat H)^*{\cal E}_{\overline L}
\end{eqnarray*}
hold. It follows that the Euler--Lagrange equations for $L$ and the Hamilton
equations for $H$ (\ref{cmp8}) are equivalent.
\end{rem}
We will say that a set of Hamiltonian forms
$H$ associated with a semiregular Lagrangian $L$ is
complete if, for each
solution
$s$ of the Euler-Lagrange equations, there exists a solution
$r$ of the Hamilton equations for a Hamiltonian form $H$ from this
set such that $s=\pi_{\Pi Y}\circ r$.
By virtue of Theorem \ref{3.24} and Remark \ref{cmp9}, a set
of associated Hamiltonian forms
is complete if, for every solution $s$ on $X$ of the Euler-Lagrange
equations for $L$, there is a Hamiltonian form $H$ from this
set which fulfills the relation
\begin{equation}
\widehat H\circ \widehat L\circ J^1s=J^1s. \label{2.36'}
\end{equation}
As for the existence of complete sets of associated Hamiltonian forms,
we refer to the following theorem. A Lagrangian $L$ is said to be almost
regular if (i) $L$ is semiregular,
(ii) the Lagrangian constraint space $N_L$ is a closed
imbedded subbundle $i_N:N_L\hookrightarrow\Pi$ of the Legendre bundle $\Pi\to
Y$ and (iii) the Legendre map
\begin{equation}
\widehat L:J^1Y\to N_L \label{cmp12}
\end{equation}
is a submersion, i.e.,
a fibred manifold.
\begin{prop}$^{2,23}$ Let $L$ be an almost regular
Lagrangian. On an open neighbourhood in $\Pi$ of each point $q\in N_L$, there
exist local Hamiltonian forms associated with $L$ which constitute a
complete set.
\end{prop}
In the case of an almost regular Lagrangian $L$, we can say something more on
the relations between Lagrangian and Hamiltonian systems as follows.
Let us assume that the fibred manifold (\ref{cmp12}) admits a
global section
$\Psi$. Let us consider the pull-back
\begin{equation}
H_N=\Psi^*H_L, \label{b4300}
\end{equation}
called the constrained Hamiltonian form.
By virtue of Lemma
\ref{3.22}, it is uniquely defined for all sections of the fibred manifold
$J^1Y\to N_L$, and $H_L=\widehat L^*H_N$.
For sections
$r$ of the fiber bundle $N_L\to X$, we can write the
constrained Hamilton equations
\begin{equation}
r^*(u_N\rfloor dH_N) =0 \label{N44}
\end{equation}
where $u_N$ is an arbitrary vertical vector field on $N_L\to X$. These
equations possess the following important properties.
\begin{theo} \label{cmp22} For any Hamiltonian form $H$
associated with an almost regular Lagrangian $L$, every solution $r$
of the Hamilton equations which lives in the Lagrangian constraint space $N_L$
is a solution of the constrained Hamilton equations (\ref{N44}).
\end{theo}
\begin{proof} Such a Hamiltonian form $H$ defines
the global section $\Psi=\widehat H\circ i_N$ of the fibred manifold (\ref{cmp12}).
Due to the relation (\ref{2.32}), $H_N=i^*_NH$ and the constrained Hamilton
equations can be written as
\begin{equation}
r^*(u_N\rfloor di^*_NH)=r^*(u_N\rfloor dH\mid_{N_L}) =0. \label{N44'}
\end{equation}
Note that they differ from the Hamilton equations
(\ref{N7}) restricted to $N_L$ which read
\begin{equation}
r^*(u\rfloor dH\mid_{N_L}) =0, \label{cmp10}
\end{equation}
where $r$ is a section of $N_L\to X$ and $u$ is an arbitrary vertical vector
field on $\Pi\to X$. A solution $r$ of the equations
(\ref{cmp10}) satisfies obviously the weaker condition
(\ref{N44'}).
\end{proof}
\begin{theo}\label{3.02}
The constrained Hamilton equations (\ref{N44}) are equivalent to the
Hamilton--De Donder equations (\ref{N46}).
\end{theo}
\begin{proof}
In accordance with the relation (\ref{LZP}), the projection $\pi_{Z\Pi}$
(\ref{b418'}) yields a surjection of $Z_L$ onto $N_L$. Given a section
$\Psi$ of the fibred manifold (\ref{cmp12}), we have the morphism
\begin{eqnarray*}
\widehat H_L\circ \Psi: N_L\to Z_L.
\end{eqnarray*}
By virtue of Lemma (\ref{3.22}), this is a surjection such that
\begin{eqnarray*}
\pi_{Z\Pi}\circ\widehat H_L\circ \Psi={\rm Id\,}_{N_L}.
\end{eqnarray*}
Hence, $\widehat H_L\circ \Psi$ is a bundle isomorphism over $Y$ which is
independent of the choice of a global section $\Psi$. Combining (\ref{cmp14})
and (\ref{b4300}) gives
\begin{eqnarray*}
H_N=(\widehat H_L\circ \Psi)^*\Xi_L
\end{eqnarray*}
that leads to the desired equivalence.
\end{proof}
The above proof gives more. Namely, since $Z_L$ and $N_L$ are isomorphic,
the Legendre morphism $H_L$ fulfills
the conditions of Theorem \ref{ddd}. Then combining Theorem \ref{ddd} and
Theorem \ref{3.02}, we obtain the following theorem.
\begin{theo}\label{3.01} Let $L$ be an almost regular Lagrangian
such that the fibred manifold (\ref{cmp12}) has a global section. A section
$\overline s$ of the jet bundle
$J^1Y\to X$ is a solution of the Cartan equations (\ref{C28}) iff
$\widehat L\circ
\overline s$ is a solution of the constrained Hamilton equations (\ref{N44}).
\end{theo}
Theorem \ref{3.01} is also a corollary of the following Lemma \ref{cmp84}.
The constrained Hamiltonian form $H_N$ (\ref{b4300}) defines the
constrained Lagrangian
\begin{equation}
L_N=h_0(H_N)=(J^1i_N)^*L_H \label{cmp81}
\end{equation}
on the jet manifold $J^1N_L$ of the fiber bundle $N_L\to X$.
\begin{lem} \label{cmp84}
There are the relations
\begin{equation}
\overline L=(J^1\widehat L)^*L_N, \qquad L_N=(J^1\Psi)^*\overline L, \label{cmp85}
\end{equation}
where $\overline L$ is the Lagrangian (\ref{cmp80}).
\end{lem}
\begin{proof}
The first of the relations (\ref{cmp85}) is an immediate consequence of the
relation (\ref{cmp83}). The latter follows from the expression (\ref{cmp5})
and the relation (\ref{b481'}) if we put $\Psi=\widehat H\circ i_N$ for some
Hamiltonian form $H$ associated with the almost regular Lagrangian $L$.
\end{proof}
The Euler--Lagrange equation for the constrained Lagrangian $L_N$ (\ref{cmp81})
are equivalent to the constrained Hamilton equations and, by virtue of Lemma
\ref{cmp84}, are quasi-equivalent to the Cartan equations. At the same
time, Cartan equations of degenerate Lagrangian systems contain an additional
freedom in comparison with the restricted Hamilton equations (see the next
Section).
The correspondence between Lagrangian and covariant Hamiltonian
dynamics of classical fields can be extended to symmetry currents and
conservation laws as follows.$^{3,21}$
Given a projectable vector
field $u$ on a fiber bundle $Y\to X$ and its lift
\begin{equation}
\widetilde u=u^\mu\partial_\mu + u^i\partial_i +( - \partial_i u^j p^\lambda_j -\partial_\mu u^\mu p^\lambda_i
+\partial_\mu u^\lambda p^\mu_i)\partial^i_\lambda \label{cmp90}
\end{equation}
onto the Legendre bundle $\Pi$, we obtain
\begin{equation}
{\bf L}_{\widetilde u}H= {\bf L}_{J^1\widetilde u}L_H, \label{b4180}
\end{equation}
i.e., the Hamiltonian form $H$ (\ref{b418}) and the
Lagrangian $L_H$ (\ref{Q3}) have the same symmetries. If the Lie derivatives
(\ref{b4180}) vanish, the corresponding symmetry current
\begin{equation}
J_u = h_0(\widetilde u\rfloor H)\mid_{y^i_\mu=\partial^i_\mu{\cal H}} \label{991}
\end{equation}
is conserved on-shell. In particular, if $u$ is a vertical vector field, we
have
\begin{equation}
J_u =\widetilde u\rfloor H=u\rfloor H. \label{cmp129}
\end{equation}
\begin{prop} Let a
Hamiltonian form $H$ be associated
with a semiregular Lagrangian $L$. Let $r$ be a solution
of the Hamilton equations for $H$ which lives
in the Lagrangian constraint space $N_L$. Let $s=\pi_{\Pi Y}\circ r$ be the
corresponding solution of the Euler--Lagrange equations for
$L$ so that the relation (\ref{2.36'}) holds. Then, for any
projectable vector field $u$ on the fiber bundle $Y\to X$, we have
\begin{eqnarray*}
J_u(r)={\cal T}_u(\widehat H\circ r),\qquad
J_u(\widehat L\circ J^1s) ={\cal T}_u(s),
\end{eqnarray*}
where ${\cal T}$ is the current (\ref{Q30}) on $J^1Y$ and $J$ is the
current (\ref{991}) on $\Pi$.
\end{prop}
It follows that the constrained Hamilton equations have symmetries of the
Euler--Lagrange equations. At the same time, given a Hamiltonian form $H$
associated with a degenerate Lagrangian $L$, the Lagrangian $L_H$ (\ref{Q3})
contains gauge fixing terms in comparison with $L$
and the Lagrangian $L_N$ (\ref{cmp81}) (see the next Section).
\bigskip
\noindent
{\bf VI. QUADRATIC DEGENERATE SYSTEMS}
\bigskip
This Section is devoted to the physically important case of almost regular
quadratic Lagrangians.
Given a fiber bundle $Y\to X$,
let us consider a quadratic Lagrangian $L$ which has the coordinate
expression
\begin{equation}
{\cal L}=\frac12 a^{\lambda\mu}_{ij} y^i_\lambda y^j_\mu +
b^\lambda_i y^i_\lambda + c, \label{N12}
\end{equation}
where $a$, $b$ and $c$ are local functions on $Y$. This property is
coordinate-independent due to the affine transformation law of the coordinates
$y^i_\lambda$. The associated Legendre map
\begin{equation}
p^\lambda_i\circ\widehat L= a^{\lambda\mu}_{ij} y^j_\mu +b^\lambda_i \label{N13}
\end{equation}
is an affine morphism over $Y$. It defines the corresponding linear
morphism
\begin{equation}
\overline L: T^*X\op\otimes_YVY\op\to_Y\Pi,\qquad p^\lambda_i\circ\overline
L=a^{\lambda\mu}_{ij}\overline y^j_\mu, \label{N13'}
\end{equation}
where $\overline y^j_\mu$ are bundle coordinates on the vector
bundle $T^*X\op\otimes_Y VY$.
\begin{lem} The Lagrangian (\ref{N12}) is semiregular.
\end{lem}
\begin{proof} Solutions $y^i_\mu$ of the pointwise linear algebraic
equations (\ref{N13}) form an affine space
modelled over the linear space of solutions of the linear
algebraic equations $a^{\lambda\mu}_{ij} \overline y^j_\mu=0$.
At each point of $N_L$, these spaces are obviously connected.
\end{proof}
Let the Lagrangian $L$ (\ref{N12}) be almost regular, i.e.,
the matrix function $a^{\lambda\mu}_{ij}$ is of constant rank. Then
the Lagrangian constraint space $N_L$
(\ref{N13}) is an affine subbundle of the Legendre bundle $\Pi\to Y$, modelled
over the vector subbundle $\overline N_L$ (\ref{N13'}) of $\Pi\to
Y$.
Hence, $N_L\to Y$ has a global section. For the sake of simplicity, let us
assume that it is the canonical
zero section $\widehat 0(Y)$ of $\Pi\to Y$. Then $\overline N_L=N_L$.
Accordingly, the kernel
of the Legendre map (\ref{N13}) is an affine
subbundle of the affine jet bundle $J^1Y\to Y$, modelled over the kernel of
the linear morphism $\overline L$ (\ref{N13'}). Then there exists a connection
\begin{eqnarray}
&&\Gamma: Y\to {\rm Ker\,}\widehat L\subset J^1Y, \label{N16}\\
&& a^{\lambda\mu}_{ij}\Gamma^j_\mu + b^\lambda_i =0, \label{250}
\end{eqnarray}
on $Y\to X$.
Connections (\ref{N16}) constitute an affine space modelled over the linear
space of soldering forms $\phi$ on $Y\to X$ satisfying the conditions
\begin{equation}
a^{\lambda\mu}_{ij}\phi^j_\mu =0 \label{cmp21}
\end{equation}
and, as a consequence, the conditions $\phi^i_\lambda b^\lambda_i=0$.
If the Lagrangian (\ref{N12}) is regular, the
connection (\ref{N16}) is unique.
\begin{lem}\label{04.2} There exists a linear bundle
map
\begin{equation}
\sigma: \Pi\op\to_Y T^*X\op\otimes_YVY, \qquad
\overline y^i_\lambda\circ\sigma =\sigma^{ij}_{\lambda\mu}p^\mu_j, \label{N17}
\end{equation}
such that $\overline L\circ\sigma\circ i_N= i_N$.
\end{lem}
\begin{proof}
The map (\ref{N17}) is a solution of the algebraic equations
\begin{equation}
a^{\lambda\mu}_{ij}\sigma^{jk}_{\mu\alpha}a^{\alpha\nu}_{kb}=a^{\lambda\nu}_{ib}.
\label{N45}
\end{equation}
After pointwise diagonalization, the matrix
$a$ has some non-vanishing components $a^{AA}$, $A\in I$. Then a
solution of the equations (\ref{N45}) takes the form
\begin{eqnarray*}
\sigma_{AA}=(a^{AA})^{-1}, \quad \sigma_{AA'}=0, \quad A\neq A',\quad A,A'\in
I,
\end{eqnarray*}
while the remaining components $\sigma_{BC}$, $B\not\in I$, are arbitrary.
In particular, there is a solution with
\begin{equation}
\sigma_{BC}=0, \qquad B\not\in I.
\label{b4120}
\end{equation}
It satisfies the particular relation
\begin{equation}
\sigma=\sigma\circ\overline L\circ\sigma. \label{N21}
\end{equation}
Further on, we will take $\sigma$ to be the solution (\ref{b4120}).
If the Lagrangian (\ref{N12}) is regular, the linear map
(\ref{N17}) is uniquely determined by the equations (\ref{N45}).
\end{proof}
The following theorem is the key point of our consideration.
\begin{theo}
There are the splittings
\begin{eqalph}
&& J^1Y={\cal S}(J^1Y)\op\oplus_Y {\cal F}(J^1Y)={\rm Ker\,}\widehat L\op\oplus_Y{\rm Im}(\sigma\circ
\widehat L), \label{N18} \\
&& y^i_\lambda={\cal S}^i_\lambda+{\cal F}^i_\lambda= [y^i_\lambda
-\sigma^{ik}_{\lambda\alpha} (a^{\alpha\mu}_{kj}y^j_\mu + b^\alpha_k)]+
[\sigma^{ik}_{\lambda\alpha} (a^{\alpha\mu}_{kj}y^j_\mu + b^\alpha_k)], \label{b4122}
\end{eqalph}
\begin{eqalph}
&& \Pi={\cal R}(\Pi)\op\oplus_Y{\cal P}(\Pi)={\rm Ker\,}\sigma \op\oplus_Y N_L, \label{N20} \\
&& p^\lambda_i = {\cal R}^\lambda_i+{\cal P}^\lambda_i= [p^\lambda_i -
a^{\lambda\mu}_{ij}\sigma^{jk}_{\mu\alpha}p^\alpha_k] +
[a^{\lambda\mu}_{ij}\sigma^{jk}_{\mu\alpha}p^\alpha_k]. \label{N20'}
\end{eqalph}
\end{theo}
\begin{proof} The proof follows from a direct computation by means of the
relations (\ref{250}), (\ref{N45}) and (\ref{N21}).
\end{proof}
It is readily observed that, with respect to the coordinates ${\cal S}^i_\lambda$
and ${\cal F}^i_\lambda$ (\ref{b4122}), the Lagrangian (\ref{N12}) reads
\begin{equation}
{\cal L}=\frac12 a^{\lambda\mu}_{ij}{\cal F}^i_\lambda{\cal F}^j_\mu +c'. \label{cmp31}
\end{equation}
Note that, in gauge theory, we have the canonical splitting (\ref{N18}) where
$2{\cal F}$ is the strength tensor.$^{3,8,9}$ The
Yang--Mills Lagrangian of gauge theory is exactly of the form (\ref{cmp31})
where $c'=0$. The Lagrangian of Proca fields is also of the form (\ref{cmp31})
where
$c'$ is the mass term. This is an example of a degenerate
Lagrangian system without gauge symmetries.
Given the linear map $\sigma$ (\ref{N17}) and a connection $\Gamma$
(\ref{N16}), let us consider the affine Hamiltonian map
\begin{equation}
\Phi=\widehat\Gamma+\sigma:\Pi \op\to J^1Y, \qquad
\Phi^i_\lambda = \Gamma^i_\lambda + \sigma^{ij}_{\lambda\mu}p^\mu_j, \label{N19}
\end{equation}
and the Hamiltonian form
\begin{eqnarray}
&& H=H_\Phi +\Phi^*L= p^\lambda_idy^i\wedge\omega_\lambda - [\Gamma^i_\lambda
(p^\lambda_i-\frac12 b^\lambda_i) +\frac12 \sigma^{ij}_{\lambda\mu}p^\lambda_ip^\mu_j-c]\omega=
\label{N22}\\
&& \qquad ({\cal R}^\lambda_i+{\cal P}^\lambda_i)dy^i\wedge\omega_\lambda - [({\cal R}^\lambda_i+{\cal P}^\lambda_i)\Gamma^i_\lambda
+\frac12
\sigma^{ij}_{\lambda\mu}{\cal P}^\lambda_i{\cal P}^\mu_j-c']\omega.\nonumber
\end{eqnarray}
\begin{theo} \label{cmp30} The Hamiltonian forms (\ref{N22})
spanned by connections $\Gamma$ (\ref{N16}) are associated with the
Lagrangian (\ref{N12}) and constitute a complete set.
\end{theo}
\begin{proof}
By the very definitions of $\Gamma$ and $\sigma$, the Hamiltonian map (\ref{N19})
satisfies the condition (\ref{2.30a}). A direct computation shows that
$\Phi=\widehat H$. Then the relation (\ref{2.30b}) also holds and, if
$\Gamma$ is a connection (\ref{N16}), the Hamiltonian
form $H$ (\ref{N22}) is associated with the Lagrangian (\ref{N12}).
Let us write the corresponding Hamilton equations (\ref{b4100a}) for
a section $r$ of the Legendre bundle $\Pi\to X$. They are
\begin{equation}
J^1s= (\widehat\Gamma+\sigma)\circ r, \qquad s=\pi_{\Pi Y}\circ r. \label{N29}
\end{equation}
Due to the surjections ${\cal S}$ and ${\cal F}$ (\ref{N18}),
the Hamilton equations (\ref{N29}) break in two parts
\begin{eqnarray}
&&{\cal S}\circ J^1s=\Gamma\circ s, \label{N23}\\
&&\partial_\lambda r^i-
\sigma^{ik}_{\lambda\alpha} (a^{\alpha\mu}_{kj}\partial_\mu r^j + b^\alpha_k)=\Gamma^i_\lambda\circ s,
\nonumber \\
&&{\cal F} \circ J^1s=\sigma\circ r, \label{N28}\\
&&\sigma^{ik}_{\lambda\alpha} (a^{\alpha\mu}_{kj}\partial_\mu r^j + b^\alpha_k)=
\sigma^{ik}_{\lambda\alpha}r^\alpha_k.\nonumber
\end{eqnarray}
Let $s$ be an arbitrary section of $Y\to X$,
e.g., a solution of the Euler--Lagrange
equations. There exists a connection $\Gamma$ (\ref{N16}) such
that the relation (\ref{N23}) holds, namely, $\Gamma={\cal S}\circ\Gamma'$ where
$\Gamma'$ is a
connection on $Y\to X$ which has $s$ as an integral section.
It is easily seen that, in this case, the Hamiltonian map (\ref{N19})
satisfies the relation (\ref{2.36'}) for $s$.
Hence, the Hamiltonian forms (\ref{N22}) constitute
a complete set.
\end{proof}
Of course, this complete set is neither minimal nor
unique. Hamiltonian forms $H$ (\ref{N22}) of this set differ from
each other in the term $\phi^i_\lambda{\cal R}^\lambda_i$ where $\phi$ are the soldering forms
(\ref{cmp21}). If follows from the splitting (\ref{N20}) that this term
vanishes on the Lagrangian constraint space.
The corresponding constrained Hamiltonian form
$H_N=i_N^*H$ and the constrained Hamilton equations (\ref{N44}) can be
written. In the case of quadratic Lagrangians, we can improve Theorem
\ref{cmp22} as follows.
\begin{theo} \label{cmp23} For every Hamiltonian
form $H$ (\ref{N22}),
the Hamilton equations (\ref{b4100b}) and (\ref{N28}) restricted to the
Lagrangian constraint space $N_L$ are equivalent to the constrained Hamilton
equations.
\end{theo}
\begin{proof} Due to the splitting (\ref{N20}), we have the corresponding
splitting
of the vertical tangent bundle $V_Y\Pi$ of the Legendre bundle $\Pi\to Y$.
In particular, any
vertical vector field
$u$ on
$\Pi\to X$ admits the decomposition
\begin{eqnarray*}
&& u= [u-u_{TN}] + u_{TN}, \\
&& u_{TN}=u^i\partial_i +a^{\lambda\mu}_{ij}\sigma^{jk}_{\mu\alpha}u^\alpha_k\partial_\lambda^i,
\end{eqnarray*}
such that $u_N=u_{TN}\mid_{N_L}$ is a vertical vector field on the Lagrangian
constraint space $N_L\to X$. Let us consider the equations
\begin{equation}
r^*(u_{TN}\rfloor dH)=0 \label{cmp15}
\end{equation}
where $r$ is a section of $\Pi\to X$ and $u$ is an arbitrary vertical vector
field on $\Pi\to X$. They are equivalent to the pair of equations
\begin{eqalph}
&& r^*(a^{\lambda\mu}_{ij}\sigma^{jk}_{\mu\alpha}\partial_\lambda^i\rfloor dH)=0,
\label{b4125a} \\
&& r^*(\partial_i\rfloor dH)=0. \label{b4125b}
\end{eqalph}
The equations (\ref{b4125b}) are obviously the Hamilton equations
(\ref{b4100b}) for $H$. Bearing in mind the relations (\ref{250}) and
(\ref{N21}), one can easily show that the equations (\ref{b4125a})
coincide with the Hamilton equations (\ref{N28}). The proof is
completed by observing that, restricted to the Lagrangian constraint space
$N_L$, the equations (\ref{cmp15}) are exactly the constrained Hamilton
equations (\ref{N44'}).
\end{proof}
Note that, in Hamiltonian gauge theory, the restricted Hamiltonian form and
the restricted Hamilton equations are gauge invariant.
Theorem \ref{cmp23} shows that, restricted to the Lagrangian constraint
space, the Hamilton equations for different Hamiltonian forms (\ref{N22})
associated with the same quadratic Lagrangian (\ref{N12}) differ from each
other in the equations (\ref{N23}). These equations are independent of
momenta and play the role of gauge-type conditions as follows.
By virtue of Theorem \ref{3.01}, the constrained Hamilton equation are
quasi-equivalent to the Cartan equations. A section
$\overline s$ of
$J^1Y\to X$ is a solution of the Cartan equations for an almost
regular quadratic Lagrangian (\ref{N12}) iff
$r=\widehat L\circ \overline s$ is a solution of the
Hamilton equations (\ref{b4100b}) and (\ref{N28}).
In particular,
let $\overline s$ be such a solution of the Cartan
equations and
$\overline s_0$ a section of the fiber bundle
$T^*X\op\otimes_Y VY\to X$ which takes its values into ${\rm Ker\,} \overline L$ (see
(\ref{N13'})) and projects onto the section
$s=\pi^1_0\circ \overline s$ of $Y\to X$. Then the affine sum $\overline s +\overline s_0$ over
$s(X)\subset Y$ is also a solution of the Cartan equations.
Thus, we come to the notion of a gauge-type freedom of the Cartan equations for
an almost regular quadratic Lagrangian $L$. One
can speak of the gauge classes of solutions of the Cartan equations whose
elements differ from each other in the above-mentioned sections $\overline s_0$. Let
$z$ be such a gauge class whose elements project onto a section $s$ of $Y\to
X$. For different connections $\Gamma$ (\ref{N16}), we consider the condition
\begin{equation}
{\cal S}\circ\overline s=\Gamma\circ s, \qquad \overline s\in z. \label{cmp25}
\end{equation}
\begin{prop} \label{cmp26}
(i) If two elements $\overline s$ and $\overline s'$ of the same gauge class
$z$ obey the same condition (\ref{cmp25}), then $\overline s=\overline s'$.
(ii) For any solution $\overline s$ of the Cartan equations, there exists a
connection (\ref{N16}) which fulfills the condition (\ref{cmp25}).
\end{prop}
\begin{proof}
(i) Let us consider the affine difference $\overline s-\overline s'$ over
$s(X)\subset Y$. We have ${\cal S}(\overline s-\overline s')=0$ iff $\overline s=\overline s'$.
(ii) In the proof of Theorem \ref{cmp30}, we have shown that, given
$s=\pi^0_1\circ\overline s$, there exists a connection
$\Gamma$ (\ref{N16}) which fulfills the relation (\ref{N23}). Let us consider the
affine difference ${\cal S}(\overline s- J^1s)$ over $s(X)\subset Y$. This is a local
section of the vector bundle ${\rm Ker\,} \overline L\to Y$ over $s(X)$. Let $\phi$ be its
prolongation onto $Y$. It is easy to see that $\Gamma+\phi$ is the
desired connection.
\end{proof}
Due to the properties in Proposition \ref{cmp26}, one can treat
(\ref{cmp25}) as a gauge-type condition on solutions
of the Cartan equations. The Hamilton equations (\ref{N23}) exemplify this
gauge-type condition when
$\overline s= J^1s$ is a solution of the Euler--Lagrange equations. At the same
time, the above-mentioned freedom characterizes solutions of the Cartan
equations, but not the Euler--Lagrange ones. First of all, this freedom
reflects the degeneracy of the Cartan equations (\ref{b336a}). Therefore, in
the Hamiltonian gauge theory, the above freedom is not related directly to the
familiar gauge invariance. Nevertheless, the Hamilton equations (\ref{N23})
are not gauge invariant, and also can play the role of gauge conditions in
gauge theory. Indeed, given a Hamiltonian form $H$ (\ref{N22}), the
corresponding Lagrangian $L_H$ (\ref{Q3}) reads
\begin{equation}
{\cal L}_H={\cal R}^\lambda_i({\cal S}^i_\lambda-\Gamma^i_\lambda) +{\cal P}^\lambda_i{\cal F}^\lambda_i
-\frac12\sigma^{ij}_{\lambda\mu}{\cal P}^\lambda_i{\cal P}^\mu_j +c'. \label{cmp86}
\end{equation}
In comparison with the Lagrangian $L$ (\ref{N12}) and the constrained
Lagrangian $L_H\mid_{J^1N_L}$,
the Lagrangian (\ref{cmp86}) includes the additional gauge fixing term
${\cal R}^\lambda_i({\cal S}^i_\lambda-\Gamma^i_\lambda)$.
\bigskip
\noindent
{\bf VII. VERTICAL EXTENSION OF POLYSYMPLECTIC FORMALISM}
\bigskip
The extension of polysymplectic
formalism to the vertical tangent bundle $VY$ of $Y\to X$ is a preliminary
step toward its BRST extension.
The Legendre bundle (\ref{00}) over $VY\to X$, called the vertical Legendre
bundle, is
\begin{eqnarray*}
\Pi_{VY}=V^*VY\op\wedge_{VY}(\op\wedge^{n-1} T^*X).
\end{eqnarray*}
We will use the compact notation
\begin{eqnarray*}
\dot\partial_i=\frac{\partial}{\partial\dot y^i}, \qquad \dot\partial^i_\lambda=\frac{\partial}{\partial\dot
p_i^\lambda}, \qquad \partial_V=\dot y^i\partial_i +
\dot p^\lambda_i \partial_\lambda^i.
\end{eqnarray*}
\begin{lem}
There exists the bundle isomorphism
\begin{equation}
\Pi_{VY}\op\cong_{VY} V\Pi,
\qquad p^\lambda_i\longleftrightarrow\dot p^\lambda_i, \qquad q^\lambda_i\longleftrightarrow p^\lambda_i, \label{cmp42}
\end{equation}
written relative to the holonomic coordinates $(x^\lambda, y^i, \dot y^i,
p^\lambda_i, q^\lambda_i)$ on $\Pi_{VY}$ and $(x^\lambda, y^i, p^\lambda_i,
\dot y^i,\dot p^\lambda_i)$ on $V\Pi$.
\end{lem}
\begin{proof}
Similar to the well-known isomorphism between the fiber bundles $TT^*X$ and
$T^*TX$,$^5$ the isomorphism
\begin{eqnarray*}
VV^*Y \op\cong_{VY} V^*VY, \quad p_i\longleftrightarrow\dot v_i,
\quad \dot p_i\longleftrightarrow\dot y_i,
\end{eqnarray*}
can be established by inspection of the
transformation laws of the holonomic
coordinates $(x^\lambda, y^i, p_i)$ on
$V^*Y$ and $(x^\lambda, y^i, v^i)$ on $VY$.
\end{proof}
It follows that
Hamiltonian formalism on the vertical Legendre bundle $\Pi_{VY}$ can be
developed as the vertical extension onto
$V\Pi$ of Hamiltonian formalism on $\Pi$, where
the canonical conjugate pairs are
$(y^i,\dot p^\lambda_i)$ and $(\dot y^i,p_i^\lambda)$. In particular,
due to the isomorphism (\ref{cmp42}),
$V\Pi$ is endowed with the canonical polysymplectic form
(\ref{406}) which reads
\begin{equation}
\Omega_{VY}=[d\dot p^\lambda_i\wedge dy^i +dp^\lambda_i\wedge d\dot y^i]\wedge\omega\otimes\partial_\lambda.
\label{cmp35}
\end{equation}
Let $Z_{VY}$ be the homogeneous Legendre bundle (\ref{N41}) over $VY$
with the corresponding coordinates $(x^\lambda,y^i,\dot
y^i,p_i^\lambda,q_i^\lambda,p)$. It can be endowed with
the canonical form $\Xi_{VY}$ (\ref{N43}). Sections of the
affine bundle
\begin{equation}
Z_{VY}\to V\Pi, \label{cmp41}
\end{equation}
by definition, provide Hamiltonian forms on $V\Pi$. Let us
consider the following particular case of these forms which are
related to those on the Legendre bundle $\Pi$.
Due to the fiber bundle
\begin{eqnarray}
&&\zeta: VZ_Y\to Z_{VY}, \label{cmp40}\\
&& (x^\lambda,y^i,\dot y^i,p_i^\lambda,q_i^\lambda,p) \circ\zeta= (x^\lambda,y^i,\dot
y^i,\dot p_i^\lambda, p_i^\lambda,\dot p), \nonumber
\end{eqnarray}
the vertical tangent bundle $VZ_Y$ of $Z_Y\to X$ is provided with
the exterior form
\begin{eqnarray*}
\Xi_V=\zeta^*\Xi_{VY}= \dot p\omega + (\dot p^\lambda_i
dy^i + p^\lambda_id\dot y^i) \wedge\omega_\lambda.
\end{eqnarray*}
Given the affine bundle
$Z_Y\to\Pi$ (\ref{b418'}), we have the fiber bundle
\begin{equation}
V\pi_{Z\Pi}: VZ_Y\to V\Pi, \label{cmp34}
\end{equation}
where $V\pi_{Z\Pi}$ is the vertical tangent map to $\pi_{Z\Pi}$. The fiber
bundles (\ref{cmp41}), (\ref{cmp40}) and (\ref{cmp34}) form the commutative
diagram.
Let $h$ be a section of the affine bundle $Z_Y\to \Pi$ and
$H=h^*\Xi$ the corresponding Hamiltonian form (\ref{b418}) on $\Pi$.
Then the section $Vh$ of the fiber bundle (\ref{cmp34}) and the
corresponding section $\zeta\circ Vh$ of the affine bundle (\ref{cmp41})
defines the Hamiltonian form
\begin{eqnarray}
&& H_V=(Vh)^*\Xi_V =(\dot p^\lambda_idy^i + p^\lambda_i d\dot
y^i)\wedge\omega_\lambda -{\cal H}_V\omega,
\label{m17}\\
&& {\cal H}_V =\partial_V{\cal H}=(\dot y^i\partial_i +\dot
p^\lambda_i\partial^i_\lambda){\cal H}, \nonumber
\end{eqnarray}
on $V\Pi$. It is called the vertical extension of $H$.
In particular, given the splitting (\ref{4.7}) of $H$ with respect to
a connection
$\Gamma$ on $Y\to X$, we have the corresponding splitting
\begin{eqnarray*}
{\cal H}_V=\dot p^\lambda_i\Gamma^i_\lambda +\dot y^j p^\lambda_i\partial_j\Gamma^i_\lambda +\partial_V\widetilde{\cal H}_\Gamma
\end{eqnarray*}
of $H_V$ with respect to the vertical connection $V\Gamma$ (\ref{cmp43}) on $VY\to
X$.
\begin{prop}
Let $\gamma$ (\ref{cmp33}) be a Hamiltonian connection on $\Pi$
associated with a Hamiltonian form $H$.
Then its vertical prolongation $V\gamma$
(\ref{cmp43}) on $V\Pi\to X$ is a Hamiltonian connection associated
with the vertical Hamiltonian form $H_V$ (\ref{m17}).
\end{prop}
\begin{proof}
The proof follows from a direct computation. We have
\begin{eqnarray*}
V\gamma=\gamma + dx^\mu\otimes [\partial_V\gamma^i_\mu\dot\partial_i +\partial_V\gamma^\lambda_{\mu i}\dot\partial_\lambda^i].
\end{eqnarray*}
Components of this connection obey the Hamilton equations (\ref{3.10})
and the equations
\begin{equation}
\dot \gamma^i_\mu=\partial^i_\mu{\cal H}_V=\partial_V\partial^i_\mu{\cal H},\qquad
\dot \gamma^\lambda_{\lambda i}=-\partial_i{\cal H}_V=-\partial_V\partial_i{\cal H}. \label{cmp51}
\end{equation}
\end{proof}
In order to clarify the physical meaning of
the Hamilton equations (\ref{cmp51}), let us suppose that
$Y\to X$ is a vector bundle. Given a solution $r$ of the
Hamilton equations for $H$, let
$\overline r$ be a Jacobi field, i.e., $r+\varepsilon \overline r$
is also a solution of the same Hamilton equations modulo terms of order $>1$ in
$\varepsilon$. Then it is readily observed that the Jacobi field $\overline r$ satisfies
the Hamilton equations (\ref{cmp51}). At the same time, the
Lagrangian $L_{H_V}$ (\ref{Q3}) on $J^1V\Pi$, defined by the Hamiltonian form
$H_V$ (\ref{m17}), takes the form
\begin{equation}
{\cal L}_{VH}=h_0(H_V)=\dot p^\lambda_i(y^i_\lambda- \partial^i_\lambda{\cal H}) -\dot y^i(p^\lambda_{\lambda
i} + \partial_i{\cal H}) +d_\lambda(p^\lambda_i\dot y^i), \label{cmp105}
\end{equation}
where $\dot p^\lambda_i$, $\dot y^i$ play the role of Lagrange multipliers. The
corresponding generating functional reduces to Dirac's
$\delta$-functions at classical solutions.
\bigskip
\noindent
{\bf VIII. BRST-EXTENDED HAMILTONIAN FORMALISL}
\bigskip
The BRST extension of Hamiltonian mechanics$^{11,12}$ shows
that: (i) one should consider vector bundles $Y\to X$ in order to introduce
generators of BRST and anti-BRST transformations, and (ii) one can narrow the
class of superfunctions under consideration because the BRST extension
of a Hamiltonian is a polynomial of a finite degree in odd variables.
Therefore, we will formulate the BRST extension on the polysymplectic
Hamiltonian formalism in the terms of simple graded manifolds.
Recall$^{24,25}$ that by a graded manifold is meant the pair $(Z,{\cal A})$ of a
smooth manifold $Z$ and a sheaf ${\cal A}$ of graded-commutative ${\bf R}$-algebras such
that
(i) there is the exact sequence of sheaves
\begin{equation}
0\to {\cal J} \to{\cal A} \to C^\infty(Z)\to 0, \qquad
{\cal J}={\cal A}_1+({\cal A}_1)^2,\label{cmp140}
\end{equation}
(ii) ${\cal J}/{\cal J}^2$ is a locally free
$C^\infty(Z)$-module of finite rank, and ${\cal A}$ is locally isomorphic to the
exterior bundle $\op\wedge_{C^\infty(Z)}({\cal J}/{\cal J}^2)$. The exact sequence
(\ref{cmp140}) admits the canonical splitting $C^\infty(Z)\to{\cal A}$, and the
well-known Batchelor's theorem takes place.
\begin{theo}$^{25,26}$
Let $(Z,{\cal A})$ be a graded manifold. There exists a vector bundle
$E\to Z$ with an $m$-dimensional
typical fiber $V$ such that ${\cal A}$ is isomorphic
to the sheaf ${\cal A}_E$ of sections of the exterior bundle
\begin{equation}
\wedge E^*={\bf R}\op\oplus_Z(\op\oplus_{k=1}^m\op\wedge^k E^*) \label{z780}
\end{equation}
whose typical fiber is the finite Grassman algebra $\wedge V^*$.
\end{theo}
This isomorphism fails to be canonical, and restricts
transformations of a graded manifold to those induced by the bundle
automorphisms of $E\to Z$. Nevertheless, this class of transformations is
sufficient for our purposes because we consider the graded extension of
Hamiltonian formalism on smooth manifolds when the vector bundle $E$
(\ref{cmp68}) below is fixed. We will call $(Z,{\cal A}_E)$ the simple graded
manifold. This is not the terminology of Ref. \cite{cari97} where this term
is applied to all finite graded manifolds, but in connection with
Batchelor's isomorphism.
Global sections of the exterior bundle (\ref{z780}) are called
superfunctions due to the equivalence between the graded manifolds $(Z,{\cal A})$
and the De Witt supermanifolds whose body is $Z$.$^{25,28}$ This isomorphism
is important for for functional integration over superfunctions. Superfunctions
make up a
${\bf Z}_2$-graded ring ${\cal A}_E(X)$.
Let $\{c^a\}$ be the holonomic bases for $E^*\to Z$ with respect to some bundle
atlas with transition functions $\{\rho^a_b\}$, i.e.,
$c'^a=\rho^a_b(z)c^b$. Then superfunctions read
\begin{equation}
f=\op\sum_{k=0}^m \frac1{k!}f_{a_1\ldots
a_k}c^{a_1}\cdots c^{a_k}, \label{z785}
\end{equation}
where $f_{a_1\cdots
a_k}$ are local functions on $Z$, and we omit the symbol of exterior product of
elements $c$. The coordinate transformation law of superfunctions
(\ref{z785}) is obvious. We will use the
notation $\nw .$ of the Grassman parity.
Given a graded manifold $(Z,{\cal A})$, the sheaf $\rm Der{\cal A}$ of graded derivations
of ${\cal A}$ is introduced. This is a subsheaf of endomorphisms of ${\cal A}$ whose
sections
$u$ on an open subset $U\subset Z$ are graded derivations of the restriction
${\cal A}\mid_U$ of the sheaf ${\cal A}$ to $U$, i.e.,
\begin{eqnarray*}
u(ff')=u(f)f'+(-1)^{\nw u\nw f}fu (f')
\end{eqnarray*}
for the homogeneous elements $u\in(\rm Der{\cal A})(U)$ and $f,f'\in {\cal A}\mid_U$.
In the case of graded manifolds, derivations of ${\cal A}$ are local operators. It
means that $(\rm Der{\cal A})(U)=\rm Der{\cal A}(U)$, i.e., if $U'\subset U$ are open
sets, there is the restriction morphism $\rm Der{\cal A}(U)\to
\rm Der{\cal A}(U')$.$^{25}$ It follows that the sheaf $\rm Der{\cal A}$ coincides with the
sheaf of graded ${\cal A}$ modules $U\to\rm Der{\cal A}(U)$.
Its sections are
called supervector fields on a manifold $Z$.
The dual of the sheaf $\rm Der{\cal A}$ is the sheaf
$\rm Der^*{\cal A}$ generated by the ${\cal A}$-linear morphisms
\begin{equation}
\phi:\rm Der{\cal A}(U)\to {\cal A}_U. \label{z789}
\end{equation}
One can think of its sections as being 1-superforms on a manifold $Z$.
In the case of a simple graded manifold $(Z,{\cal A}_E)$ supervector fields and
1-superforms can be represented by sections of vector bundles as follows.
Due to the canonical splitting
$VE= E\times E$, the vertical tangent bundle
$VE\to E$ can be provided with the fiber bases $\{\partial_a\}$ dual of $\{c^a\}$.
These are fiber bases for ${\rm pr}_2VE=E$. Let $(z^A)$ be coordinates on $Z$. Then
a supervector field on a trivialization domain $U$ read
$u= u^A\partial_A + u^a\partial_a$
where $u^A, u^a$ are local superfunctions.
It yields a graded endomorphism of ${\cal A}_E(U)$ by the rule
\begin{equation}
u(f_{a\ldots b}c^a\cdots c^b)=u^A\partial_A(f_{a\ldots b})c^a\cdots c^b +u^a
f_{a\ldots b}\partial_a\rfloor (c^a\cdots c^b). \label{cmp50'}
\end{equation}
This implies the corresponding
coordinate transformation law
\begin{eqnarray*}
u'^A =u^A, \qquad u'^a=\rho^a_ju^j +u^A\partial_A(\rho^a_j)c^j
\end{eqnarray*}
of supervector fields. It follows that supervector fields on $Z$, which we
agree to call $E$-determined supervector fields, can be represented by
sections of the vector bundle
${\cal V}_E\to Z$ which is locally isomorphic to the vector bundle
\begin{eqnarray*}
{\cal V}_E\mid_U\approx\wedge E^*\op\otimes_Z({\rm pr}_2VE\op\oplus_Z TZ)\mid_U,
\end{eqnarray*}
and has the transition functions
\begin{eqnarray*}
&& z'^A_{i_1\ldots i_k}=\rho^{-1}{}_{i_1}^{a_1}\cdots
\rho^{-1}{}_{i_k}^{a_k} z^A_{a_1\ldots a_k}, \\
&& v'^i_{j_1\ldots j_k}=\rho^{-1}{}_{j_1}^{b_1}\cdots
\rho^{-1}{}_{j_k}^{b_k}\left[\rho^i_jv^j_{b_1\ldots b_k}+ \frac{k!}{(k-1)!}
z^A_{b_1\ldots b_{k-1}}\partial_A(\rho^i_{b_k})\right]
\end{eqnarray*}
of the bundle coordinates $(z^A_{a_1\ldots a_k},v^i_{b_1\ldots b_k})$,
$k=0,\ldots,m$. These transition functions
fulfill the cocycle relations.
There is the exact sequence over $Z$ of vector
bundles
\begin{equation}
0\to \wedge E^*\op\otimes_Z{\rm pr}_2VE\to{\cal V}_E\to \wedge E^*\op\otimes_Z TZ\to 0. \label{cmp92}
\end{equation}
Due to the above mentioned locality property the
sheaf of sections of the vector bundle
${\cal V}_E\to Z$ is isomorphic to the sheaf $\rm Der\,{\cal A}_E$. Global sections of
${\cal V}_E\to Z$ constitute the ${\cal A}_E(Z)$-module of supervector fields on $Z$,
which is also a Lie
superalgebra with respect to the bracket
\begin{eqnarray*}
[u,u']=uu' + (-1)^{\nw u\nw{u'}+1}u'u.
\end{eqnarray*}
One can think of a splitting
\begin{equation}
\widetilde\gamma:\dot z^A\partial_A \mapsto \dot z^A(\partial_A +\widetilde\gamma_A^a\partial_a) \label{cmp70}
\end{equation}
of the exact sequence (\ref{cmp92}) as being a graded connection, though
this is not a true connection on ${\cal V}_E\to Z$. A
graded connection can be represented by a section
\begin{equation}
\widetilde \gamma=dz^A\otimes(\partial_A +\widetilde\gamma^a_A\partial_a) \label{cmp93}
\end{equation}
of the vector bundle $T^*Z\op\otimes_Z{\cal V}_E\to Z$ such that the composition
\begin{eqnarray*}
Z\op\to^{\widetilde\gamma}T^*Z\op\otimes_Z{\cal V}_E\to T^*Z\op\otimes_Z (\wedge
E^*\op\otimes_Z TZ)\to T^*Z\op\otimes_ZTZ
\end{eqnarray*}
is the canonical form $dz^A\otimes\partial_A$ on $Z$.
Such a graded connection $\widetilde\gamma$ transforms every vector field $\tau$ on $Z$
into a supervector field
\begin{eqnarray*}
\tau=\tau^A\partial_a\mapsto \widetilde\gamma\tau=\tau^A(\partial_A +\widetilde\gamma_A^a\partial_a),
\end{eqnarray*}
and provides the corresponding decomposition
\begin{eqnarray*}
u= u^A\partial_A + u^a\partial_a=u^A(\partial_A +\widetilde\gamma_A^a\partial_a) + (u^a-
u^A\widetilde\gamma_A^a)\partial_a
\end{eqnarray*}
of supervector fields on $Z$.
For instance, every linear connection
\begin{eqnarray*}
\gamma=dz^A\otimes (\partial_A +\gamma_A{}^a{}_bv^b\partial_a)
\end{eqnarray*}
on the vector bundle $E\to Z$ defines the graded connection
\begin{equation}
\gamma_S=dz^A\otimes (\partial_A +\gamma_A{}^a{}_bc^b\partial_a) \label{cmp73}
\end{equation}
such that, for any vector field $\tau$ on $Z$ and any superfunction $f$,
the graded derivation $\gamma_S\tau(f)$ is exactly the covariant derivative
$\tau^A\nabla_A f$ relative to the connection $\gamma$.
\begin{rem}
Let now $Z\to X$ be a fiber bundle, coordinated by $(x^\lambda,z^i)$.
Let
\begin{eqnarray*}
\gamma=\Gamma +\gamma_\lambda{}^a{}_bv^bdx^\lambda\otimes\partial_a
\end{eqnarray*}
be a connection on $E\to X$ which is a linear morphism over a connection $\Gamma$
on $Z\to X$. Then we have the bundle monomorphism
\begin{eqnarray*}
\gamma_S: \wedge E^*\op\otimes_ZTX\ni u^\lambda\partial_\lambda \mapsto u^\lambda(\partial_\lambda+\Gamma^i_\lambda\partial_i
+\gamma_\lambda{}^a{}_bc^b\partial_a)\in {\cal V}_E
\end{eqnarray*}
over $Z$, called a composite graded connection on $Z\to X$.
It is represented by a section
\begin{equation}
\gamma_S= \Gamma + \gamma_\lambda{}^a{}_bc^bdx^\lambda\otimes\partial_a \label{cmp122}
\end{equation}
of the fiber bundle $T^*X\op\otimes_Z{\cal V}_E\to Z$ such that the composition
\begin{eqnarray*}
Z\op\to^{\gamma_S}T^*X\op\otimes_Z{\cal V}_E\to T^*X\op\otimes_Z (\wedge
E^*\op\otimes_Z TZ)\to T^*X\op\otimes_ZTX
\end{eqnarray*}
is the pull-back onto $Z$ of the canonical form $dx^\lambda\otimes\partial_\lambda$ on $X$.
\end{rem}
The $\wedge E^*$-dual ${\cal V}^*_E$ of ${\cal V}_E$ is a vector bundle over $Z$
which is locally isomorphic to the vector bundle
\begin{eqnarray*}
{\cal V}^*_E\mid_U\approx \wedge E^*\op\otimes_Z({\rm pr}_2VE^*\op\oplus_Z T^*Z)\mid_U,
\end{eqnarray*}
and has the transition functions
\begin{eqnarray*}
&& v'_{j_1\ldots j_kj}= \rho^{-1}{}_{j_1}^{a_1}\cdots
\rho^{-1}{}_{j_k}^{a_k} \rho^{-1}{}_j^a v_{a_1\ldots a_ka}, \nonumber\\
&& z'_{i_1\ldots i_kA}=
\rho^{-1}{}_{i_1}^{b_1}\cdots
\rho^{-1}{}_{i_k}^{b_k}\left[z_{b_1\ldots b_kA}+ \frac{k!}{(k-1)!}
v_{b_1\ldots b_kj}\partial_A(\rho^j_{b_k})\right]
\end{eqnarray*}
of the bundle coordinates $(z_{a_1\ldots a_kA},v_{b_1\ldots b_kj})$,
$k=0,\ldots,m$, with respect to the dual bases $\{dz^A\}$ for $T^*Z$ and
$\{dc^b\}$ for ${\rm pr}_2V^*E=E^*$. There is the exact sequence
\begin{equation}
0\to \wedge E^*\op\otimes_ZT^*Z\to{\cal V}^*_E\to \wedge E^*\op\otimes_Z {\rm pr}_2VE^*\to 0.
\label{cmp72}
\end{equation}
The sheaf of sections of ${\cal V}^*_E\to Z$ is isomorphic to
the sheaf $\rm Der^*{\cal A}_E$. Global sections of the vector bundle
${\cal V}^*\to Z$ constitute the ${\cal A}_E(Z)$-module of $E$-determined exterior
1-superforms $\phi=\phi_A dz^A + \phi_adc^a$
on $Z$ with the coordinate transformation law
\begin{eqnarray*}
\phi'_a=\rho^{-1}{}_a^b\phi_b, \qquad \phi'_A=\phi_A
+\rho^{-1}{}_a^b\partial_A(\rho^a_j)\phi_bc^j.
\end{eqnarray*}
Then
the morphism (\ref{z789}) can be seen as the interior product
\begin{equation}
u\rfloor \phi=u^A\phi_A + (-1)^{\nw{\phi_a}}u^a\phi_a. \label{cmp65}
\end{equation}
Any graded connection $\widetilde\gamma$ (\ref{cmp93}) also yields the
splitting of the exact sequence (\ref{cmp72}), and defines the corresponding
decomposition of 1-superforms
\begin{eqnarray*}
\phi=\phi_A dz^A + \phi_adc^a =(\phi_A+\phi_a\widetilde\gamma_A^a)dz^A +\phi_a(dc^a
-\widetilde\gamma_A^adz^A).
\end{eqnarray*}
Accordingly, $k$-superforms $\phi$ are sections
of the graded exterior bundle $\overline\wedge^k_Z{\cal V}^*_E$ such that
\begin{eqnarray*}
\phi\overline\wedge\sigma =(-1)^{\nm\phi\nm\sigma +\nw\phi\nw\sigma}\sigma\overline\wedge \phi.
\end{eqnarray*}
The interior product (\ref{cmp65})
is extended to higher degree superforms by the rule
\begin{eqnarray*}
u\rfloor (\phi\overline\wedge\sigma)=(u\rfloor \phi)\overline\wedge \sigma
+(-1)^{\nm\phi+\nw\phi\nw{u}}\phi\overline\wedge(u\rfloor\sigma).
\end{eqnarray*}
Recall that the graded exterior differential
$d$ of superfunctions is introduced in accordance with the condition
$u\rfloor df=u(f)$
for an arbitrary supervector field $u$, and is
extended uniquely to higher degree superforms by the rules
\begin{eqnarray*}
d(\phi\overline\wedge\sigma)= (d\phi)\overline\wedge\sigma +(-1)^{\nm\phi}\phi\overline\wedge(d\sigma), \qquad d\circ d=0.
\end{eqnarray*}
It takes the coordinate form
\begin{eqnarray*}
d\phi= dz^A \overline\wedge \partial_A(\phi) +dc^a\overline\wedge \partial_a(\phi),
\end{eqnarray*}
where the left derivatives
$\partial_A$, $\partial_a$ act on the coefficients of superforms by the rule
(\ref{cmp50'}), and they are graded commutative with the forms $dz^A$, $dc^a$.
The Lie
derivative of a superform $\phi$ along a supervector field $u$ is given by
the familiar formula
\begin{equation}
{\bf L}_u\phi= u\rfloor d\phi + d(u\rfloor\phi). \label{cmp66}
\end{equation}
\begin{rem} \label{cmp102}
Given a vector bundle $E\to Z$, let us consider the jet manifold $J^1E$,
coordinated by $(z^A,v^a,v^a_A)$. This is also a vector bundle over $Z$. Then
one can construct the corresponding fiber bundles ${\cal V}_{J^1E}$ and
${\cal V}_{J^1E}^*$. Due to the monomorphism $E^*\to (J^1E)^*$, there is
the monomorphism
${\cal V}_E^*\to
{\cal V}_{J^1E}^*$, i.e., every
$E$-determined superform on $Z$ can be also seen as a $J^1E$-determined
superform. In particular, the horizontal projection
$h_0$ (\ref{cmp100}) gives rise to the 0-graded homomorphism
\begin{equation}
h_0: dc^a\to c^a_Adz^A \label{cmp101}
\end{equation}
which sends $E$-determined superforms onto horizontal $J^1E$-determined
superforms.
\end{rem}
Turn now to the BRST extension of covariant Hamiltonian formalism on the
Legendre bundle $\Pi$ (\ref{00}) when $Y\to X$ is a vector bundle. Let us
apply the above construction of simple graded manifolds to the case of the
vertical tangent bundle
\begin{equation}
E=VV\Pi=V\Pi\op\oplus_X V\Pi\op\longrightarrow^{{\rm pr}_1} V\Pi \label{cmp68}
\end{equation}
over the vertical Legendre bundle $Z=V\Pi\to X$.
Let $(x^\lambda, y^i, p^\lambda_i,\dot y^i,\dot p^\lambda_i)$ be the holonomic
coordinates on $V\Pi$. Then the dual $E^*$ of $E$ can be endowed with the
associated fiber bases
$\{c^i,c_i^\lambda,\overline c^i,\overline c_i^\lambda\}$ such that
$c^i$ and $\overline c^i$ have the same linear coordinate
transformation law as the coordinates $y^i$ and $\dot y^i$, while
$c_i^\lambda$ and $\overline c_i^\lambda$ have
those of the coordinates $p_i^\lambda$ and $\dot p_i^\lambda$. The corresponding
supervector fields and superforms are introduced on $V\Pi$ as sections of the
vector bundles ${\cal V}_{VV\Pi}$ and ${\cal V}^*_{VV\Pi}$, respectively. Let us
complexify these bundles as ${\bf C}\op\otimes_X{\cal V}_{VV\Pi}$ and
${\bf C}\op\otimes_X{\cal V}^*_{VV\Pi}$.
As in mechanics, the main criterion of the BRST extension of covariant
Hamiltonian formalism is its invariance under BRST and anti-BRST
transformations whose generators are the supervector fields
\begin{eqnarray}
&& \vartheta_Q= \partial_c +i\dot y^i
\frac{\partial}{\partial \overline c^i} + i\dot p_i^\lambda\frac{\partial}{\partial\overline c_i^\lambda},
\quad \vartheta_{\overline Q}=\partial_{\overline c} -i\dot y^i
\frac{\partial}{\partial c^i} - i\dot p_i^\lambda\frac{\partial}{\partial c_i^\lambda}, \label{z746}\\
&& \partial_c=c^i\partial_i + c_i^\lambda\partial^i_\lambda, \qquad
\partial_{\overline c}=\overline c^i\partial_i + \overline c_i^\lambda\partial^i_\lambda, \nonumber
\end{eqnarray}
on $V\Pi$. They fulfill the nilpotency rules
\begin{eqnarray*}
\vartheta_Q\vartheta_Q=0, \qquad \vartheta_{\overline Q} \vartheta_{\overline Q}=0, \qquad \vartheta_{\overline Q} \vartheta_Q
+\vartheta_Q\vartheta_{\overline Q}=0.
\end{eqnarray*}
The BRST- and
anti-BRST-invariant extension of the polysymplectic form
$\Omega_{VY}$ (\ref{cmp35}) on $V\Pi$ is the
$TX$-valued superform
\begin{eqnarray*}
\Omega_S=[d\dot p_i^\lambda\wedge dy^i +dp_i^\lambda\wedge d\dot y^i +i(d\overline c_i^\lambda\wedge dc^i- d\overline
c^i\wedge dc_i^\lambda)]\wedge \omega\otimes\partial_\lambda
\end{eqnarray*}
on $V\Pi$, where $(c^i,-i\overline c^\lambda_i)$ and $(\overline c^i,i c^\lambda_i)$ are the
conjugate pairs.
Let $\gamma$ be a Hamiltonian connection for a Hamiltonian form $H$. The double
vertical connection $VV\gamma$ on $VV\Pi\to X$ is a linear morphism over the
vertical connection $V\gamma$ on $V\Pi\to X$, and so defines the composite
graded connection
\begin{eqnarray*}
(VV\gamma)_S =V\gamma + dx^\mu\otimes[\overline g^i_\mu\frac{\partial}{\partial \overline c^i} +\overline
g^\lambda_{\mu i}\frac{\partial}{\partial\overline c_i^\lambda} + g^i_\mu\frac{\partial}{\partial c^i} +g^\lambda_{\mu
i}\frac{\partial}{\partial c_i^\lambda}]
\end{eqnarray*}
(\ref{cmp122}) on $V\Pi\to X$, whose components $g$ and $\overline g$ are given by
the expressions
\begin{eqnarray*}
\overline g^i_\lambda=\partial_{\overline c}\partial^i_\lambda{\cal H}, \quad
\overline g^\lambda_{\lambda i}=-\partial_{\overline c}\partial_i{\cal H},\quad
g^i_\lambda= \partial_c\partial^i_\lambda{\cal H},
\quad g^\lambda_{\lambda i}= -\partial_c\partial_i{\cal H}.
\end{eqnarray*}
This composite graded connection satisfies the relation
\begin{eqnarray*}
(VV\gamma)_S\rfloor\Omega_S=-dH_S,
\end{eqnarray*}
and so is a Hamiltonian graded connection for the Hamiltonian
superform
\begin{equation}
H_S=[\dot p_i^\lambda dy^i +p_i^\lambda d\dot y^i +i(\overline c_i^\lambda dc^i +d\overline
c^i c_i^\lambda)]\omega_\lambda -(\partial_V
+ i\partial_{\overline c}\partial_c){\cal H}\omega \label{z750}
\end{equation}
on $V\Pi$. This superform is BRST- and anti-BRST-invariant, i.e., ${\bf L}_\vartheta
H_S=0$. Thus, it is the desired BRST extension of the Hamiltonian form $H$.
The Hamiltonian superform $H_S$ (\ref{z750}) defines the
corresponding BRST-extension of the Lagrangian $L_H$ (\ref{Q3}). Following
Remark \ref{cmp102}, let us consider the vector bundle
\begin{eqnarray*}
J^1(VV\Pi)=VJ^1(V\Pi)\to J^1(V\Pi)=VJ^1\Pi
\end{eqnarray*}
and the corresponding fiber bundle ${\cal V}_{VJ^1(V\Pi)}\to VJ^1\Pi$. It is
readily observed that $VV\Pi$-determined superforms on $V\Pi$ can be seen as
particular $VJ^1(V\Pi)$-determined superforms on $VJ^1\Pi$. Moreover,
combining the horizontal projections $h_0$ (\ref{cmp100}) and (\ref{cmp101})
for exterior forms and superforms, we obtain the 0-graded homomorphism
$h_0$ which sends $VV\Pi$-determined superforms on $V\Pi$ onto the horizontal
$VJ^1(V\Pi)$-determined superforms on $VJ^1\Pi\to X$. Then the horizontal
superdensity
\begin{eqnarray}
&& L_{SH}=h_0(H_S)= L_{VH} + i[(\overline c_i^\lambda
c^i_\lambda+\overline c^i_\lambda c_i^\lambda) -\partial_{\overline c}\partial_c{\cal H}]\omega= L_{VH}
+\label{cmp104}\\
&& \qquad i[\overline c^\lambda_i(c^i_\lambda -\partial_c\partial^i_\lambda{\cal H}) + (\overline
c^i_\lambda -\partial_{\overline c}\partial^i_\lambda{\cal H})c^\lambda_i + \overline
c^\lambda_ic^\mu_j\partial^i_\lambda\partial^j_\mu{\cal H} -
\overline c^ic^j\partial_i\partial_j{\cal H}]\omega
\nonumber
\end{eqnarray}
on $VJ^1\Pi\to X$ can be treated as the desired BRST extension of the
Lagrangian $L_H$ (\ref{Q3}). Note that, in
comparison with the Lagrangian $L_{VH}$ (\ref{cmp105}), the generating
functional determined by the BRST-extended Lagrangian
(\ref{cmp104}) is not reduced to $\delta$-functions.
The BRST-extended Lagrangian $L_{SH}$ (\ref{cmp104}) is also invariant under
the jet prolongations
\begin{eqnarray*}
J^1\vartheta=\vartheta^a\partial_a+ d_\lambda\vartheta^a\partial_a^\lambda
\end{eqnarray*}
of the BRST and anti-BRST transformations (\ref{z746}).
Moreover, it is easily verified that both
$L_{HS}$ and
$H_S$ are invariant under transformations whose generators are the
supervector fields
\begin{eqnarray}
&& \vartheta_K=c_i^\lambda\frac{\partial}{\partial \overline c_i^\lambda} +c^i\frac{\partial}{\partial \overline c^i}, \quad
\vartheta_{\overline K}=\overline c_i^\lambda\frac{\partial}{\partial c_i^\lambda} +\overline c^i\frac{\partial}{\partial c^i},
\label{cmp130}\\
&& \vartheta_C=c_i^\lambda\frac{\partial}{\partial c_i^\lambda} +c^i\frac{\partial}{\partial c^i} -
\overline c_i^\lambda\frac{\partial}{\partial \overline c_i^\lambda} -\overline c^i\frac{\partial}{\partial\overline c^i}.
\nonumber
\end{eqnarray}
The supervector fields (\ref{z746}) and (\ref{cmp130}) constitute the
Lie superalgebra of the well-known group ISp(2):
\begin{eqnarray}
&& [Q,Q]=[\overline Q,\overline Q]=[\overline Q, Q]=[K,Q]=[\overline K,\overline Q]=0, \label{cmp128} \\
&& [K,\overline Q]=Q, \quad [\overline K, Q]=\overline Q, \quad [K,\overline K]= C, \quad
[C,K]=2K, \quad [C,\overline K]=-2\overline K. \nonumber
\end{eqnarray}
Similarly to the lift $\widetilde u$ (\ref{cmp90}) onto $\Pi$ of a vector field $u$
on $Y$, the supervector fields $\vartheta$ (\ref{z746}) and (\ref{cmp130}) can be
represented as the corresponding graded lift
\begin{eqnarray*}
\vartheta=\widetilde u=u^a\partial_a - (-1)^{[y^a]([p_b] +[u^b])}\partial_au^b\frac{\partial}{\partial p_a}
\end{eqnarray*}
of some $VVY$-determined supervector fields $u$ on $VY$ which are sections
of the fiber bundle ${\cal V}_{VVY\to VY}$. These supervector fields $u$ read
\begin{eqnarray}
&& u_Q= c^i\partial_i + i\dot y^i
\frac{\partial}{\partial \overline c^i},
\quad u_{\overline Q}=\overline c^i\partial_i -i\dot y^i
\frac{\partial}{\partial c^i}, \label{cmp131} \\
&& u_K=c_i^\lambda\frac{\partial}{\partial \overline c_i^\lambda}, \quad
u_{\overline K}=\overline c_i^\lambda\frac{\partial}{\partial c_i^\lambda},
\quad u_C=c^i\frac{\partial}{\partial c^i} -\overline c^i\frac{\partial}{\partial\overline c^i}.
\nonumber
\end{eqnarray}
They also constitute the Lie superalgebra (\ref{cmp128}).
Then by analogy with (\ref{cmp129}), we obtain the corresponding
supercurrents
$J_\vartheta=\vartheta\rfloor H_S=u\rfloor H_S$. These are the horizontal
$(n-1)$-superforms
\begin{eqnarray*}
&& Q=(c^i\dot p^\lambda_i-\dot y^i c_i^\lambda)\omega_\lambda, \quad \overline Q=(\overline c^i\dot
p^\lambda_i-\dot y^i \overline c_i^\lambda)\omega_\lambda, \\
&& K=-i c_i^\lambda c^i\omega_\lambda, \quad \overline K=i\overline c_i^\lambda \overline c^i\omega_\lambda,\quad
C=i(\overline c^\lambda c^i -\overline c^i c_i)\omega_\lambda
\end{eqnarray*}
on $V\Pi$. They form the Lie superalgebra (\ref{cmp128}) with respect to the
product (\ref{c2}). It should be emphasized that
the Lie superalgebra (\ref{cmp128}) provides the canonical symmetries of
any BRST-extended Hamiltonian system.
The following construction is similar to that is met in supersymmetric
mechanics and BRST mechanics.
Given a function $F$ on the Legendre bundle $\Pi$,
let us consider the operators
\begin{equation}
F_\beta=e^{\beta F}\circ\vartheta\circ e^{-\beta F} =\vartheta
-\beta\partial_c F, \quad
\overline F_\beta=e^{-\beta F}\circ\overline\vartheta\circ e^{\beta F}=\overline\vartheta
+\beta\partial_{\overline c} F, \quad \beta>0, \label{cmp135}
\end{equation}
called the BRST and anti-BRST charges, which act on superfunctions
on $V\Pi$. These operators are nilpotent, i.e.,
\begin{equation}
F_\beta\circ F_\beta=0, \qquad \overline F_\beta\circ \overline F_\beta=0. \label{z763}
\end{equation}
By the BRST- and anti-BRST-invariant extension of $F$ is meant the
superfunction
\begin{equation}
F_S= -\frac{i}{\beta}(\overline F_\beta\circ F_\beta +F_\beta\circ \overline F_\beta).
\label{cmp133}
\end{equation}
We have the relations
\begin{eqnarray*}
F_\beta\circ F_S- F_S\circ F_\beta=0, \qquad
\overline F_\beta\circ F_S- F_S\circ\overline F_\beta=0.
\end{eqnarray*}
These relations together with the relations (\ref{z763}) provide the
operators
$F_\beta$, $\overline F_\beta$, and $F_S$ with the structure of the Lie
superalgebra sl(1/1).$^{29}$
Let now $\Gamma^i_\lambda=\Gamma_\lambda{}^i{}_jy^j$ be a linear
connection on $Y\to X$ and $\widetilde\Gamma$ some Hamiltonian connection on
$\Pi\to X$ for the Hamiltonian form $H_\Gamma$ (\ref{3.6}).
Given the splitting
(\ref{4.7}) of the Hamiltonian form
$H$ with respect to the connection $\Gamma$, there is the corresponding splitting
of the BRST-extended Hamiltonian form
\begin{eqnarray*}
{\cal H}_S={\cal H}_{\Gamma S} +\widetilde{\cal H}_{S\Gamma}=\dot p_i^\lambda\Gamma_\lambda{}^i{}_jy^j +\dot
y^jp_i^\lambda\Gamma_\lambda{}^i{}_j +i(\overline c_i^\lambda\Gamma_\lambda{}^i{}_jc^j +\overline
c^j\Gamma_\lambda{}^i{}_jc_i^\lambda) +(\partial_V + i\partial_{\overline c}\partial_c)\widetilde{\cal H}_\Gamma
\end{eqnarray*}
with respect to the composite graded connection $(VV\widetilde\Gamma)_S$
(\ref{cmp122}) on the fiber bundle $V\Pi\to X$. Let $dV$ be a volume element
on $X$ and $\widetilde{\cal H}_\Gamma\omega=FdV$, where $F$ is a function on $\Pi$. Then
\begin{eqnarray*}
\widetilde{\cal H}_{S\Gamma}\omega = F_SdV=-i(\overline F_1\circ F_1 +F_1\circ \overline F_1)dV,
\end{eqnarray*}
where $F_1$, $\overline F_1$ and $F_S$ are the BRST and anti-BRST charges
(\ref{cmp135}) and (\ref{cmp133}). The similar splitting of a
super-Hamiltonian is the corner stone of supersymmetric mechanics.$^{30,31}$
|
\section{Introduction}
Radio and infrared observations have revealed that stars are formed
in dense cloud cores (e.g., \cite{beich86,myers87}).
From these observations as well as theoretical works,
models of protostellar evolution for low-mass stars,
from embedded protostars to visible T Tauri stars,
have been proposed (e.g., Shu, Adams, \& Lizano 1987).
Recently, even kinematic evidence for the infall of dense
cores around embedded sources is observed (e.g., Zhou et al. 1993;
Hayashi, Ohashi, \& Miyama 1993). Although our understanding of the embedded
and T Tauri phases has made great progress, the earliest stage of
star formation or even a prior stage to star formation has still
been poorly understood.
Dense cores without any detectable young stellar objects,
i.e., starless dense cores are most probably sites where star
formation has just started or will soon start.
Although some dense cores, such as B 335 (Zhou et al. 1990),
have been found to be a very early collapse phase,
even in this case significant material has already collapsed
onto the central star. Starless cores are therefore good targets
to study the earliest stage of star formation.
L1544 is a well-studied starless core in Taurus, observed in
several line emissions such as NH$_3$, CS, and N$_2$H$^+$
(e.g., \cite{bm89,tafalla98}; hereafter T98,
Williams et al. 1999; hereafter W99) and
submillimeter continuum (Ward-Thompson et al. 1994).
Their observations were, however, made mainly using single-dish telescopes,
which prevented them from studying detailed geometrical and kinematical
structures of L1544.
In addition, most of the emission lines they used were optically thick,
resulting in difficulty in
investigating the whole velocity structure without suffering
self-absorption.
In this Letter, we report results of
interferometric observations of the L1544 starless core in CCS ($J_N=3_2-2_1$).
This transition of CCS will be excited in cold cores such as
starless cores because of its low
energy level ($E_u\sim3.2$~K; \cite{yamamoto90}).
Interferometric observations enable us to study
geometrical and kinematical structures of L1544 in detail.
In addition, CCS in L1544 is optically thin (see \S 2).
Moreover, CCS is one of the carbon chain molecules that are more abundant
in starless cores (\cite{suzuki92}),
and have no hyperfine structures, meaning that it is a good probe
to study geometrical and kinematical structures of starless cores in detail
(\cite{langer95}, Kuiper, Langer, \& Velusamy 1996, \cite{wolkovitch97}).
\section{Observations}
Observations were made using the nine-antenna
Berkeley-Illinois-Maryland Association (BIMA) array
\footnote{Operated by the University of California at Berkeley,
the University of Illinois, and the University of Maryland,
with support from the National Science Foundation.}
in September 1998.
We observed CCS ($J_N=3_2-2_1$), which has rest frequency of 33.751374 GHz
(\cite{yamamoto90}), with
low-noise receivers utilizing cooled HEMT amplifiers,
developed for observations of
the Sunyaev-Zeldovich effect (\cite{carl96}).
The filed of view of the array is $\sim$5\arcmin\ at this frequency.
Such a wide field of view is of great advantage to the imaging of L1544,
which is moderately extended (see T98).
The typical system temperature during
the observations was 40-60~K in Single-side-band.
Spectral information was obtained using a digital correlator
with 1024 channels at a bandwidth of 6.25~MHz,
providing a velocity resolution of $\sim$0.054 km~s$^{-1}$\
at the observed frequency.
Two different configurations of the array were used.
Projected baselines ranged from 6~m to 240~m,
so that the observations were not sensitive to structures extended
more than $\sim$5\arcmin, corresponding to 42000~AU at the distance of Taurus
($d=140$~pc; \cite{elias78})
\footnote{Note that the observations are less sensitive to the structures
close to 5\arcmin\ in extent (see Wilner \& Welch 1994)}.
The obtained channel data was reduced using the MIRIAD package.
The phase was calibrated by observing 0530+135,
and the complex passband of each baseline was determined from
observations of 3C84. When CCS maps were made and cleaned,
a Gaussian taper was applied to the visibility data to improve
sensitivity to extended low brightness emission,
so that the resultant beam size was
20\arcsec$\times$13\arcsec\ with a position angle of $-5$\arcdeg.
The resultant 1 $\sigma$
rms noise level for channel maps was typically 0.165~Jy~beam$^{-1}$,
equivalent to $\sim$0.7~K in brightness.
In order to measure the optical depth of CCS in L1544,
we also observed CCS and CC$^{34}$S ($J_N=3_2-2_1$) simultaneously
using the Nobeyama 45~m telescope ($\Delta\theta\sim52\arcsec$) in April 1999.
$T_{\rm A}^*$ of CCS was measured to be 1.1~K, while no CC$^{34}$S emission
was detected to a 3 $\sigma$ level of 0.072~K in $T_{\rm A}^*$,
indicating that a 3 $\sigma$ upper limit to the optical depth of CCS
is 0.93 when the sulfur isotope ratio is the terrestrial value, 23.
In this Letter, we will discuss results of the BIMA observations in detail.
\section{Results}
CCS was detected at the LSR velocities
ranging from 6.88 to 7.48~km~s$^{-1}$\ with high S/N ($\geq 3\sigma$).
Figure~1 shows the CCS total intensity map, integrated over
this velocity range.
The map shows a structure elongated in the northwest-southeast
direction (PA $\sim$144\arcdeg), with a size of
$\sim$210\arcsec $\times$ 64\arcsec\
at the 3 $\sigma$ level,
corresponding to 0.15 $\times$ 0.045~pc at the distance to Taurus.
The ratio between the major and minor axes is $\sim$3.3.
The condensation consists of two blobs, one at the northwest and
the other at the southeast, and each blob contains several peaks,
suggesting that the condensation has clumpy structures.
A similar elongated structure was also observed in N$_2$H$^+$ (W99),
whereas it
shows a smaller size (7000~AU $\times$ 3000~AU) and
a slightly large position angle (PA$=$155\arcdeg).
More interestingly,
the CCS map does not show a prominent peak at the central part
of the condensation (marked by the cross in Fig.~1),
where the other maps taken in C$^{34}$S (2-1), N$_2$H$^+$ (1-0), and 800~$\mu$m \
show prominent peaks (T98, W99, \cite{wt94}).
This difference suggests that the CCS in L1544
traces the outer regions of the core while the other molecules and
dust trace the inner part.
The clumpy structures of the CCS condensation are more obvious
in the channel maps as shown in the left panels of Fig.~2.
Remarkable characteristics of the clumps are their
narrow line width: each clump is detected at only a few velocity channels.
This is shown in the right panels of Fig.~2, where
line profiles of two representative clumps are presented.
The line widths deconvolved with the instrumental velocity resolution
(0.054~km~s$^{-1}$) of the clumps
were measured to be (0.06$\pm$0.01)-(0.13$\pm$0.01)~km~s$^{-1}$\ in
full width at half-maximum,
which are almost equivalent to the thermal line width of
the CCS molecule at 10~K or 0.09~km~s$^{-1}$\ (see Langer et al. 1995),
suggesting that thermal motions are dominant in the clumps.
Note that the line profiles of the clumps often show double peaks.
This is because
two clumps with different peak LSR velocities
partially coincide with each other.
The multi-peak-velocity structure of L1544 was also observed in C$^{34}$S (T98).
Physical
properties of the clumps will be discussed in detail in a forth coming paper.
We discern the global velocity field of the CCS condensation from
position-velocity diagrams (Fig. 3). Two velocity components are
visible along the projected minor axis of the condensation
(Fig.~3a): one at
$V_{\scriptsize LSR}\sim$7.1~km~s$^{-1}$\ and the other
at $V_{\scriptsize LSR}\sim$7.4~km~s$^{-1}$.
Both components
persist all along the minor axis except at the southwestern edge, where
only the 7.4~km~s$^{-1}$\ component is detected, and at the
northeastern edge, where only the 7.1~km~s$^{-1}$\ component is detected.
This suggests that the two
velocity components are mostly coincident along the line-of-sight.
Neither velocity component shows any significant gradient along the
minor axis.
The PV diagram along the projected major axis (Fig.~3b) shows more
remarkable velocity structures.
Similar to the PV diagram along the minor axis,
there are two velocity components in the inner parts of the condensation
($-40$\arcsec$ \leq \Delta x \leq +40$\arcsec\ in Fig.~3b),
whereas these two components merged into a single velocity component
at each of the southeastern and northwestern edges of the condensation.
The velocity difference between the two velocity components
at $\Delta x=0$\arcsec\ is $\sim$0.25~km~s$^{-1}$,
similar to that in Fig.~3a.
In addition, a global velocity gradient,
$\sim$0.08~km~s$^{-1}$\ per 14000 AU, is observed from the southeast to the northwest.
The N$_2$H$^+$ observations (W99) also showed a velocity gradient at inner parts
of L1544 along almost the same direction (PA$=$155\arcdeg)
although their measurements showed a $\sim$3 times larger gradient.
Thus, the whole velocity structure of the CCS condensation
can be represented as a
\lq\lq tilted ellipse\rq\rq, as shown by the dashed curve in Fig.~3b.
The elliptical velocity structure is
roughly symmetrical with respect to $V_{\scriptsize LSR}\sim$7.25~km~s$^{-1}$,
suggesting that this velocity seems to be the systemic velocity of this system.
This suggests that the two velocity components seen in Fig.~3a
are actually blueshifted and redshifted parts
in a single kinematical system.
The global velocity gradient seen along the projected major axis
suggests rotation of the CCS condensation, while the blueshifted and
redshifted components overlapping with each other along the line of sight
can be explained by inward motions in the condensation (see \S4.2.).
\section{Discussion}
\subsection{Geometrical Structures of the CCS Condensation}
The elongated structure seen in the CCS total intensity map suggests that
the CCS condensation has either a filamentary structure or a flattened
one with an almost edge-on configuration.
The latter case is more likely
in view of the high column densities needed to
produce the self-absorption features observed in optically thick
molecular tracers such as CS (T98) or N$_2$H$^+$ (W99).
The envelopes around young stellar objects (YSOs)
often show
flattened structures similar to the CCS condensation here
except that their sizes are smaller (e.g., \cite{ohashi97}).
This fact suggests that flattened envelopes
are present even before the YSOs form.
The flattened geometry of starless cores are predicted by some theoretical
simulations (e.g., \cite{nakamura95,matsumoto97}), in which
magnetic fields or rotation play an important role in producing
flattened structures.
Note that magnetic fields may be more important to explain the flattened
geometry of CCS because of its slow rotation (see \S4.2.; see also W99).
For the L1544 CCS condensation,
the ratio between
the projected major and minor axes implies
an inclination angle of $\sim$73\arcdeg\ (0\arcdeg\ for the face-on case)
if it is spatially thin.
If the flattened condensation is spatially thick
like the envelopes around YSOs (e.g., L1527; \cite{ohashi97}),
then this estimate is a lower limit to the true inclination.
As pointed out in \S3, the CCS total intensity is
stronger at the northwestern and southeastern edges of the
condensation, and weaker in the center.
A ringlike geometry for the condensation, observed almost edge-on,
is consistent with this pattern, since, for optically thin emission, the
total intensity is proportional to the column density
as long as there is no significant gradient in the excitation.
As we will show in the next section,
the velocity field also suggests a ringlike geometry for the condensation.
Note that the CCS ring is not a physical structure but rather probably
results from a lower abundance of CCS towards the center of the L1544 core.
Most other tracers, including C$^{34}$S, N$_2$H$^+$, and submillimeter dust continuum,
show prominent peaks close to the center of the L1544 core. A similar
situation has been observed in L1498, another starless core in Taurus,
where CCS is weak toward the dust continuum peak (\cite{kuiper96,willacy98}).
The apparent structure may be due to chemical evolution.
As a dense core collapses and the density increases, the photoionization
and photodissociation processes become gradually less effective in the
central region with highest density. The abundance of molecules sensitive
to the photochemistry changes, and the CCS abundance decreases substantially
(\cite{suzuki92}). In addition, CCS depletes onto grains in a collapsing gas
that increases the density (\cite{bergin97}),
also explaining the lower abundance of CCS toward the center of L1544.
These ideas of the chemical evolution are consistent
with the result that L1544
is undergoing infall, as evidenced by kinematic tracers (see \S4.2).
\subsection{Kinematical Structures of the CCS Condensation}
As shown in \S3, the kinematics of the CCS condensation are characterized
by two major features: (1) blueshifted and redshifted components that
coincide with each other along the line of sight, and (2) a velocity
gradient along the major axis. Taking into account the flattened edge-on
structure of the CCS condensation, the blueshifted and redshifted components
may be explained by radial motion in the plane of the condensation,
while the velocity gradient may represent rotation of the condensation.
One possible radial motion in the plane of the condensation is infall,
which we favor over expansion, as discussed below. Here we consider the
kinematics of the condensation
using a simple model that consists of a spatially thin ring
with both infall and rotation.
A simple model ring was obtained by modifying the model disk used in
the study of the flattened envelope of L1527, which exhibits both infall
and rotation (\cite{ohashi97}).
For L1544, we have modified the
surface density distribution and velocity field in two ways:
(1) The surface density is constant at
$R_{\scriptsize out} > R > R_{\scriptsize in}$, and is null
at $R \leq R_{\scriptsize in}$,
where $R_{\scriptsize out}$ and $R_{\scriptsize in}$ are the outer and
inner radii of the ring,
respectively; (2) The infall velocity, $V_{\scriptsize infall}$, is constant
throughout, while the rotation velocity,
$V_{\scriptsize rotation}$,
is proportional to the radius, i.e., rigid rotation such that
$V_{\scriptsize rotation}=V_{\scriptsize rotation}^0$($R$/$R_{\scriptsize out}$),
where $V_{\scriptsize rotation}^0$ is the rotation velocity
at $R_{\scriptsize out}$.
The first modification makes the model have a ring structure.
Under the assumption that the model ring has an edge-on configuration
with respect to the observer, PV diagrams along the projected major
axis of the model ring are calculated to compare with the observed
PV diagram (Fig.~3a).
For the model calculations,
$R_{\scriptsize out}$ was fixed at 15000~AU according
to the CCS observations, while
three parameters,
$R_{\scriptsize in}$, $V_{\scriptsize infall}$,
and $V_{\scriptsize rotation}^0$ remained adjustable.
As shown in Fig.~4a,
when $R_{\scriptsize in}=7500$~AU,
$V_{\scriptsize infall}=0.12$~km~s$^{-1}$,
and $V_{\scriptsize rotation}^0=0.09$~km~s$^{-1}$,
the observed PV diagram (Fig. 3b) is well reproduced by the model:
the calculated PV diagram shows a velocity structure, represented
as a tilted ellipse, with a velocity difference at $\Delta x=0$ of
0.24~km~s$^{-1}$\ and a global velocity gradient of
0.09~km~s$^{-1}$\ per 15000 AU. Note that $2V_{\scriptsize infall}$
corresponds to the velocity
difference at $\Delta x=0$ in the
calculated diagram, while $V_{\scriptsize rotation}^0$ is equivalent
to the velocity
gradient per 15000 AU in the diagram. Hence, it is easily
understood that much smaller or larger $V_{\scriptsize infall}$
and/or $V_{\scriptsize rotation}^0$
cannot reproduce the observed velocity structures. On the other hand,
a ring structure with a larger $R_{\scriptsize in}$
is essential for the model to reproduce
the observed PV because
when
much smaller $R_{\scriptsize in}$, including the case of
$R_{\scriptsize in}=0$ (equivalent to a model
with a disk structure) is used,
prominent peaks emerge close to
$\Delta x=0$ in the calculated diagram (see Fig. 4b). This is because when
$R_{\scriptsize in}$ decreases, the total column density
through the plane of the ring
drastically increases close to $\Delta x=0$.
One might argue that expansion instead of infall can explain
the observed PV diagram. However, infall is more likely to explain
the kinematics of the CCS condensation because the mass of L1544
derived from the virial theorem
is comparable to or much smaller than
that estimated from the 800~$\mu$m dust emission:
a virial mass is estimated to be $\sim$1.7~$M_{\sun}$
using the radius of the CCS condensation (15000~AU) and the mean line
width of the CCS (0.4~km~s$^{-1}$), under the assumption of a spherical cloud
with a constant density,
while a mass of 2-6~$M_{\sun}$ was derived from 800~$\mu$m dust emission
\footnote{Note that the current CCS data is not adequate to
derive a mass because CCS may be less abundant in the inner parts
of L1544 (see \S4.1).}
(T98).
Inward motions in L1544 were also suggested from
a spectroscopic method (T98, W99).
Our result is remarkable as the observations resolve the velocity field
and show {\it direct} evidence for infall motions
in the starless core L1544.
The kinematics of L1544 revealed by the CCS imaging show differences with
the envelopes of YSOs. The estimated infall and rotation velocities
are comparable to each other at a scale of 10000~AU, unlike YSOs where
infall velocities are inferred to be 2 to 6 times larger than rotation
velocities in the outer regions. In addition, the estimated infall velocity
is much smaller in magnitude than those inferred around YSOs
(e.g., \cite{hayashi93,ohashi97,momose98}).
Although our model ring with a constant infall velocity and rigid
rotation explain the kinematics of the CCS condensation very well,
the current data do not place strong constraints on the radial
dependences of the infall and rotation velocities because the CCS
molecule traces only the outer part of the L1544 core.
W99 have estimated infall and rotation velocities
in the inner regions of L1544, where CCS emission is weak or absent,
based on the self-absorption of N$_2$H$^+$.
Their estimates of
$\sim$0.08~km~s$^{-1}$ and $\leq$ 0.14~km~s$^{-1}$ for infall and rotation, respectively,
suggests nearly constant infall and rotation throughout the L1544 core,
from scales of 10000 to 1000~AU.
The optical depth of the N$_2$H$^+$ lines may yet mask the
innermost regions of L1544, and direct measurements of the kinematics
using an optically thin tracer may be valuable for revealing the
radial dependences of the systematic motions in this starless core.
\acknowledgments
We are grateful to P. T. P. Ho, H. Masunaga, F. Nakamura, K. Saigo, S. Takano,
and S. Yamamoto for fruitful discussions. We also thank H. Maezawa for his
help during observations with the Nobeyama 45~m telescope.
\clearpage
|
\section{Introduction}
We have improving observational evidence\cite{observations} on
how structure formed on the scale of galaxies and larger from
measurements of the angular distributions of the radiation
backgrounds,
surveys of the spatial distributions and motions of the galaxies,
and observations of the evolution of galaxies and the intergalactic
medium back to redshift $z\sim 5$. It may
prove useful to
complement these advances with explorations of the options
conventional physics offers for theories of structure formation.
Here we consider the
possibility that the dark matter that is thought to
dominate the gravitational growth of structure
interacts only with itself and
gravity, the dark matter originating by gravitational
particle production, that is, as a squeezed state,
at the end of inflation.\footnote{Such dark matter particles might
be called cabots, in honour of the Cabot family of Boston, who
were said to be so highly placed in society as to speak only to
God.}
After specifying our model in Section II, we discuss the state of
the dark matter field at the end of inflation. We begin in Section~III
with the equilibrium properties of a noninteracting dark mass
scalar field
in de~Sitter spacetime. For many purposes de~Sitter spacetime is a
useful approximation to inflation, and it has the particular advantage
that the equilibrium statistical properties of the dark mass
field are well specified \cite{AV83,Starob,SY}.
However, it is also of interest to analyze the effect of
the rolling value of the Hubble parameter in a model for
the relatively late stages of inflation. This is
presented in Section~IV, following in part the early work by Kofman
and Linde \cite{KL}.
Section V deals with the constraints on the
mass of the dark matter particles and on their self-coupling
from the conditions that a single scalar field has the wanted
present mean mass density and field fluctuations that are
compatible with the large-scale anisotropy of the
thermal background. In
quintessential inflation\cite{quint}
this dark matter candidate
has an observationally interesting effective Jeans length.
We discuss in section VI issues of internal consistency of the
model. In a recent preprint
Felder, Kofman and Linde \cite{fkl} have presented a
stimulating list of potential problems with quintessential
inflation and indirectly with the noninteracting dark matter model.
We address their points in this section.
An assessment of consistency with the full
suite of observational tests requires computations that go
well beyond our exploratory discussion; we also comment on these
open issues in Section VI.
\section{Assumptions}
The dark matter is modeled as a single scalar field,
$y({\bf x},t)$, that is minimally
coupled to gravity and has the self-interaction potential
\begin{equation}
V = \lambda y^4/4 + \mu ^2y^2/2.
\label{2}
\end{equation}
The generalization to several field
components with $O(N)$ symmetry and the opposite sign of the
quadratic part of the potential is trivial and not interesting
for the present purpose.
We assume that there is a time before the end of inflation
when the Hubble parameter $H(t)$ is close enough to constant that the
dark matter field can relax to near statistical equilibrium
between the amplitude growth driven by quantum fluctuations and
the classical slow roll due to the self-interaction potential, in an
approximation to a de Sitter-invariant quantum state.
For definiteness, much of our discussion of this assumption
is based on a quartic inflaton potential,
\begin{equation}
U=\lambda _\phi\phi ^4/4,
\label{Uphi}
\end{equation}
with a self-coupling satisfying $\lambda_\phi\lesssim 10^{-14}$,
for consistency with the isotropy of the thermal cosmic
background radiation (the CBR).
Kofman and Linde \cite{KL} consider the two-field model
in Eqs~(\ref{2}) and (\ref{Uphi}) in the limit where the
evolution of $y$ is essentially classical. Following their
methods in part, we show in Section VI that, when
\begin{equation}
\lambda\gg\lambda _\phi ,
\label{lambdas}
\end{equation}
the $y$-field energy density indeed remains subdominant
during inflation, as required for internal consistency.
We show in Section~V that the
condition~(\ref{lambdas}) is within the ranges of values of
$\lambda$ and $\lambda _\phi$ allowed by the CBR.
We assume the mass parameter $\mu$ is small enough that the quartic
part of the potential dominates during inflation, that $\lambda$ is
small enough that the field may be approximated as the sum of a
classical part and a free quantum field, and that
radiative corrections to the potential in Eq.~(\ref{2}) are
negligible. The condition for the last assumption is \cite{foot1}
\begin{equation}
{3\lambda\over{64\pi^2}}\ln{1\over{\lambda}}\ll 1.
\label{2'}
\end{equation}
We use a homogeneous cosmological model in which
the field equation is
\begin{equation}
{\partial ^2y\over\partial t^2} + 3{\dot a\over a}{\partial y\over\partial t} =
{\nabla ^2 y\over a^2} - (\lambda y^2 + \mu ^2)y.
\label{3}
\end{equation}
For $\lambda y^2\ll H^2$, the oscillation frequency due to the
quartic part of the
potential is small compared to $H$, and we can use the
slow roll approximation to the field equation,
\begin{equation}
3H\partial y/\partial t = -\lambda y^3.
\label{slowroll}
\end{equation}
This neglects the mass term, which is justified for $y^2\gg
\mu^2/\lambda$.
We consider both the conventional model
for reheating and the much lower temperature at the end of
inflation implied by the quintessential inflation
model \cite{quint}. In the latter case the mass density in
interacting matter at the end of inflation is proportional to the
fourth power of the Hubble parameter, with a coefficient that
depends on the nature of the matter and its interactions.
We comment on the uncertainty in
this coefficient in Section~VI.
We write the Planck mass as $m_{\rm pl} = [3/(8\pi G)]^{1/2}$
in units where $\hbar = 1 = c$.
The Hubble parameter during inflation is $H$ and the value at the
end of inflation is $H_x$.
In section V the present value of Hubble's constant
is $H_o=100h$ km~s$^{-1}$~Mpc$^{-1}$. Numerical examples assume
$h=0.7$, density parameter
$\Omega _m=0.3$ in matter capable of clustering,
and a cosmologically flat universe.
\section{Statistics of the Matter Field in de Sitter Spacetime}
On scales larger than the de Sitter horizon, $H^{-1}$, the dynamics of
the field $y$ can be pictured as a ``random walk'' [superimposed on
the classical slow roll in Eq.~(\ref{slowroll})]
in which $y$ undergoes random steps
of rms magnitude $(\delta y)_{\rm rms}=H/2\pi$ per expansion time
$\delta t=H^{-1}$, independently in each horizon-size region. The
statistical properties of the field $y$ resulting from this random
process can be described in terms of the probability distribution
function $P(y,t)$ which satisfies the Fokker-Planck equation
\cite{AV83,Starob}
\begin{equation}
{\partial P(y,t)\over{\partial
t}}={1\over{3H}}{\partial\over{\partial y}}[V'(y)P(y,t)] +
{H^3\over{8\pi^2}}{\partial^2 P(y,t)\over{\partial y^2}} .
\label{FP}
\end{equation}
Any initial distribution, $P(y,0)$, approaches the stationary solution of
(\ref{FP}),
\begin{equation}
P_0(y)=N^{-1}\exp\left(-{8\pi^2\over{3H^4}}V(y)\right),
\label{P0}
\end{equation}
on the timescale \cite{SY}
\begin{equation}
\tau _{\rm rel} \sim \lambda ^{-1/2}H^{-1}.
\label{tau}
\end{equation}
The coefficient $N$ in (\ref{P0}) is determined from the normalization
condition $\int_{-\infty}^{\infty}P_0(y)dy = 1$.
The stationary distribution $P_0(y)$ corresponds to the de
Sitter-invariant quantum state of the field $y$. The statistical
properties of this state in the case of a quartic
potential $V(y)=\lambda y^4/4$ have been studied by Starobinsky and
Yokoyama\cite{SY} (hereafter SY).
The distribution function $P_0(y)$ is not Gaussian; in particular it
has negative excess
kurtosis \cite{SY} $\langle\phi^4\rangle/(\langle\phi^2\rangle^2)-3 \approx 0.812$.
In the following
numerical example of correlation functions of
powers of the field, we will refer to the moments
\begin{eqnarray}
\langle y^2\rangle &=& 0.1318H^2\lambda^{-1/2}, \quad
\langle y^4\rangle - \langle y^2\rangle ^2 = 0.0206H^4\lambda^{-1}, \nonumber\\
\langle y^6\rangle &=& 0.01502H^6\lambda^{-3/2}, \quad
\langle y^8\rangle - \langle y^4\rangle ^2 = 0.00577H^8\lambda^{-2}.\label{10}\\
\langle y^{12}\rangle &-& 3\langle y^4\rangle\langle y^8\rangle + 2\langle y^4\rangle ^3
= 0.001755 H^{12}\lambda ^{-3}. \nonumber
\end{eqnarray}
SY use the Fokker-Planck formalism to compute the field
correlation function $\langle y_1y_2\rangle$. Their method
and numerical results are readily adapted to get
the correlation function of powers of the field,
in particular the
correlation function of the mass density $\lambda y^4/4$.
This is done in the Appendix; we find that for a positive integer power
$n$ of the field the equal time reduced
correlation function
of $y^n$ at large separation $x_{12}$ is
\begin{equation}
c_n(x_{12}) = \langle y_1^ny_2^n\rangle - \langle y^n\rangle ^2 \propto x_{12}^{-p},
\label{11}
\end{equation}
where the power law index is
\begin{equation}
p = 0.178\lambda ^{1/2} \hbox{ (odd } n), \qquad
p = 0.579\lambda ^{1/2} \hbox{ (even } n).
\label{12}
\end{equation}
The power law in Eq.~(\ref{11}) applies at separations large
compared to the comoving coherence length
\begin{equation}
R_c = (aH)^{-1}\exp(1/p) \sim (aH)^{-1}e^{\alpha H\tau_{rel}},
\label{rc}
\end{equation}
with the appropriate value of $p$ from Eq.~(\ref{12}),
and where $\alpha$ is a number of order unity.
This expression can be interpreted to mean that in the
relaxation time $\tau_{\rm rel}$ (Eq.~\ref{tau}) comoving
positions initially separated by the Hubble length have moved
to separation $aR_c$, with significant loss of
memory of the common initial field values at the two positions.
As an alternative to the Fokker-Planck formalism, it may be convenient
to use a discrete version of the Langevin equation.
In an expansion time, $t=H^{-1}$, the value of the
field $y$
at a fixed comoving position changes by
the amount
\begin{equation}
y(j)\rightarrow y(j+1)= y(j) -\lambda y(j)^3/(3H^2) +
\iota _jH/(2\pi ).
\label{6}
\end{equation}
The integer $j$ counts successive expansion times $H^{-1}$.
The Gaussian normal random variables $\iota _j$ have zero mean and are
statistically independent, so in particular
$\langle\iota _j\iota _k\rangle =\delta _{j,k}$.
The second term on the right-hand side of Eq.~(\ref{6}), which
represents the classical force due to the potential
that tends to drive $y$ toward zero,
follows from the slow roll approximation in Eq.~(\ref{slowroll}). The last
term represents the frozen quantum fluctuations added to $y$
in the expansion time $H^{-1}$.
In the stationary de Sitter-invariant state, the mean square value of $y$
is
independent of time.
This condition applied to Eq.~(\ref{6}), and keeping first powers of
the classical drift part and second powers of the fluctuating part,
yields the mean mass density in the $y$-field,
\begin{equation}
\langle\rho _y\rangle =\lambda\langle y^4\rangle /4 = 3H^4/(32\pi ^2).
\label{8}
\end{equation}
The condition $\langle y(j)^6\rangle = \langle y(j+1)^6\rangle$ similarly yields
\begin{equation}
\langle y^8\rangle /\langle y^4\rangle ^2 = 5.
\label{variance}
\end{equation}
These relations agree with the Fokker-Planck
results given above, as expected.
Eq.~(\ref{variance}) says the standard deviation of the frozen field
mass fluctuations
is $\delta\rho /\rho = 2$, with coherence length
$R_c\sim H^{-1}\exp(1.7\lambda^{-1/2})$.
Eq.~(\ref{6})
may be applied as a numerical iteration prescription,
where the integers $j$ count iterations and the
numbers $\iota _j$ are drawn from a generator of
pseudo-Gaussian independent
random numbers with zero mean and unit variance. We use it to
calculate the two-point functions (\ref{11}) and illustrate
the different power law indices for even and odd $n$.
The computation starts with an initial
value $y(0)$ obtained after 300
iterations of Eq.~(\ref{6}) starting from $y=0$. Two time series,
$y_1(j)$ and $y_2(j)$, are computed from the same initial value
$y(0)$ and different sets of the $\iota _j$. This is a numerical
realization of the values of the $y$-field at successive expansion
times $H^{-1}$ at two fixed comoving positions that are
separated by the Hubble length at $j=0$.
A set of realizations is obtained by using
the last value of the time series $y_1$ as the initial value
for the next realization of $y_1(j)$ and $y_2(j)$.
The mean of $y_1^n(j)y_2(j)^n$
across a set of these realizations is an
estimate of the expectation value
$\langle y_1^ny_2^n\rangle$ at separation $x_{12}\propto e^j$.
Fig. 1 shows numerical results for $n=1$, 2, 3, and 4 in
Eq.~(\ref{11}). The variables are scaled to
\begin{equation}
\hat c_n = \lambda ^{n/2}H^{-2n}c_n, \qquad
\hat x_{12} = (Hax_{12})^{\sqrt\lambda},
\label{14}
\end{equation}
to remove the dependence on the parameters $H$ and $\lambda$
at zero separation (Eq.~\ref{10}) and at large separation
(Eq.~\ref{12}). The heavy lines in the figure
are the averages across realizations for two parameter choices,
$\lambda = 0.1$ and $\lambda = 0.0001$. The latter requires a much
larger number of iterations to reach a given range of values of the
correlation functions. That reduces the number of realizations
in the computation, so the numerical noise is larger.
The straight dotted lines are interpolations based on the SY results: the
intercepts at zero separation are the
one-point moments in Eq.~(\ref{10}) and the slopes are
the power law indices 0.178 and 0.579 for the scaled
variable $(Hax)^{\sqrt\lambda}$ (Eq.~\ref{12}).
Within the
fluctuations from the limited number
of realizations in the averages ($1 \times 10^8$ for $\lambda =0.1$
and $1 \times 10^7$ for $\lambda = 1\times 10^{-4}$) the
numerical estimates are
consistent with the expected scaling with $\lambda$.
Since the power-law asymptotics apply only for $x > R_c$, one could not
expect them to match with the one-point moments which correspond to
$ax\sim H^{-1}\ll aR_c$. Fig.~1 indicates, however, that the
difference from a pure power law is not dramatic: the mismatch
is no more than a factor of about $2$.
The lowest curves in Fig. 1 show the mass correlation
function $\xi (x_{12}) \propto\langle y_1^4y_2^4\rangle - \langle y^4\rangle ^2$. At large
separation the numerical result is a factor of about
two below the extrapolation
from zero separation. Since the reduced second moment of the mass
distribution is $\xi = 4$ at zero separation (Eq.~\ref{variance}),
a useful approximation to the dimensionless dark mass
correlation function is
\begin{equation}
\xi (x) = \langle\delta _1\delta _2\rangle
\sim 2(x_x/x)^\epsilon , \qquad \epsilon = 0.58\lambda ^{1/2},
\label{15}
\end{equation}
where $x_x$ is the comoving Hubble length at the
end of inflation and the density contrast is
\begin{equation}
\delta =\rho /\langle\rho\rangle - 1 = y^4/\langle y^4\rangle - 1.
\label{16}
\end{equation}
Comparison to the moments of large-scale fluctuations in
the counts of galaxies\cite{Gaz} requires higher moments of the
$y$-matter distribution; we present the example of
the three-point function. The reduced dimensionless three-point
mass correlation function is
\begin{eqnarray}
\xi _3 &=& \langle\delta _1\delta _2\delta _3\rangle \nonumber\\
&=& \langle y_1^4y_2^4y_3^4\rangle /\langle y^4\rangle ^3-
(\langle y_1^4y_2^4\rangle +\langle y_2^4y_3^4\rangle +\langle y_3^4y_1^4\rangle )/
\langle y^4\rangle ^2 + 2.
\label{17}
\end{eqnarray}
The arguments of $\xi _3$ are the lengths of the sides of the
triangle defined by the positions at which the field is evaluated.
It is shown in the Appendix
that if the three points
define an equilateral triangle with side $x$
much larger than $R_c$ then the three-point function varies with $x$ as
$\xi _3\propto \xi ^{3/2}$, where the two-point function $\xi$
(Eq.~\ref{15}) is evaluated at $x$.
The one-point moments (Eq.~\ref{10}) show that
$\xi _3/\xi ^{3/2}=4$ at zero separation.
Fig. 2 shows numerical
realizations of $\xi _3/\xi ^{3/2}$ for equilateral triangles.
The estimates of $\xi _3$ are based on sets of
three independent time series $y_1(j)$, $y_2(j)$, and $y_3(j)$
with the same initial value. The values of $\lambda$, the numbers
of realizations, and the
scaled variables are the same as for Fig. 1. The curves are
close to the SY one-point moments at small separation,
and, as also predicted by the Fokker-Planck method,
at scaled separation $\hat x\lower.5ex\hbox{$\; \buildrel > \over \sim \;$} 10$ the ratio is close to
independent of triangle size, within the considerable noise
from the limited number of realizations, at
\begin{equation}
\xi _3/\xi ^{3/2}\simeq 0.8
\label{18}
\end{equation}
This gives a useful working approximation to the normalization
of the three-point function
on scales of interest for structure formation.\footnote{The asymptotic
value of the ratio in Eq.~(\ref{18}) can be found
by a numerical computation of the eigenfunction $\Phi_2(y)$ and
the integral $C$ in Eq.~(\ref{C}) of the Appendix.
We have not attempted this computation.}
The scaling of the $n$-point mass
correlation functions with distance $x$ at fixed values of the
ratios $x_{ij}/x_{kl}$ of distances among the $n$ points
may be obtained by the following argument.
Let $\bar\delta _i=\delta\rho/\rho$ be the mass density contrast
at expansion parameter
$a=a_i$ smoothed within a window of fixed comoving size, shape,
and position, and let $\bar\delta _f$ be the density contrast
smoothed in the same window at a later time,
at expansion parameter $a_f$. A realization $\delta (t)$ of
the evolution from $\delta _i$ to $\delta _f$ is a result of the
process of freezing of quantum fluctuations and
the classical evolution
toward $\delta = 0$. If $a_f/a_i$ is large enough that the
epochs are separated by many relaxation times, the
initial value $\delta _i$ is a small perturbation to this process,
and in the lowest nontrivial order in perturbation theory the
expectation value of $\bar\delta _f$ for given initial
value $\bar\delta _i$ is linear in $\bar\delta _i$:
\begin{equation}
\langle\bar\delta _f\rangle _{\bar\delta _i}= T\bar\delta _i .
\label{transfer_eq}
\end{equation}
The transfer coefficient $T$ is a function of the window size and
shape, which we are holding fixed in comoving coordinates,
and of the expansion factor $a_f/a_i$, which we will vary.
Next we consider two
disjoint windows, both of fixed size, shape and position in the
same comoving coordinate system. If the distance between the
windows is much larger than the Hubble length at
$a_i$ the evolution of the density contrast in window~1 is
independent of the process in window~2, and it follows that
the equal time mass two-point correlation function satisfies
\begin{equation}
\langle\bar\delta _f(1)\bar\delta _f(2)\rangle =
T^2\langle\bar\delta _i(1)\bar\delta _i(2)\rangle .
\label{AA}
\end{equation}
In the de Sitter equilibrium state the two-point
mass correlation function depends only on
the proper separation $ax$, so we can rewrite
Eq.~(\ref{AA}) as
\begin{equation}
\xi (bx) = T(b)^2 \xi (x), \qquad b = a_f/a_i.
\end{equation}
This implies
\begin{equation}
\xi\propto x^{-2\nu },\qquad T\propto b^{-\nu},
\label{AB}
\end{equation}
where $\nu$ is a constant that must be positive for convergence.
That is, at large separation $x$ the correlation
function varies as a power of $x$,
as was shown more directly by our use of the SY analysis.
The $n$-point function similarly satisfies
\begin{equation}
\langle\bar\delta _f(1)\ldots\bar\delta _f(n)\rangle =
T^n\langle\bar\delta _i(1)\ldots\bar\delta _i(n)\rangle ,
\end{equation}
from which it follows that the $n$-point function scales with
the separation $x$ at fixed ratios of separations among the $n$ points
as
\begin{equation}
\xi_n\propto x^{-n\nu },
\label{BB}
\end{equation}
consistent with Eq.~(\ref{18}).
Eq.~(\ref{BB}) says the $n^{\rm th}$ central moment
of the dark mass $M$ contained in a volume of fixed shape
scales as $M_n=\langle (M-\langle M\rangle)^n\rangle\propto\sigma ^n$, where
$\sigma=(\langle (M-\langle M\rangle)^2\rangle)^{1/2}$ is the standard
deviation (and assuming the integrals over the correlation
functions converge at vanishing separation between points).
The smoothed mass fluctuations do not approach a Gaussian
as the size of the smoothing window is increased. Rather, this
is a scale-invariant fractal in which
the mass fluctuations smoothed
and referred to the standard deviation are independent of
the size of the smoothing window.
\section{The Dark Matter at the End of Inflation}
We have assumed so far that the evolution of the inflaton $\phi$ is
sufficiently slow that the field $y$ has enough time to settle
into its de Sitter-invariant ``equilibrium'' state. Here we
specify the conditions of applicability of this picture under a
specific inflaton potential, and we consider the transition to
the opposite regime, which applies toward the end of inflation,
when the evolution of $y$ is almost classical.
For the inflaton potential in Eq.~(\ref{Uphi})
an approximate solution for the expansion history through inflation is
\cite{Lindebook}
\begin{equation}
H = H_xe^{-H_xt}, \qquad H_x={4\over 3}\lambda _\phi ^{1/2}m_{\rm pl} ,
\qquad \phi _x=\left( 8/3\right) ^{1/2}m_{\rm pl} ,
\label{Hoft}
\end{equation}
with expansion parameter
\begin{equation}
\log a_x/a = H/H_x - 1.
\label{axovera}
\end{equation}
We write the comoving length scale $x$ of field fluctuations
frozen at time $t$, when the expansion and Hubble parameters
are $a$ and $H$, as
\begin{equation}
Hax=1.
\label{xoft}
\end{equation}
This comoving length reaches its minimum value, $x_x$,
at the end of inflation. In Eq.~(\ref{Hoft}) inflation ends at
$t=0$, when the Hubble parameter and inflaton
field values are $H_x$ and $\phi _x$.
In the approximation of
Eqs. (\ref{Hoft}) to (\ref{xoft})
the Hubble parameter $H$ when the field fluctuations on the
comoving length scale $x$ are freezing satisfies
\begin{equation}
\log (x/x_x) \simeq \log (H_x/H) + H/H_x - 1.
\label{simulation}
\end{equation}
A useful approximation at $a\ll a_x$ is
\begin{equation}
x/x_x \sim e^{H/H_x}.
\label{xoverxx}
\end{equation}
The relaxation time for $y$ and the evolution time for $H$
are (Eqs.~\ref{tau} and~\ref{Hoft})
\begin{equation}
\tau _{\rm rel} \sim \lambda ^{-1/2}H^{-1},\qquad \tau _H\sim H_x^{-1}.
\label{relaxation times}
\end{equation}
When $\tau _{\rm rel}\ll\tau_H$ the $y$-field relaxes to
statistical equilibrium between fluctuations and classical slow roll,
in a good approximation to the de Sitter
equilibrium state. When $\tau _{\rm rel}\gg\tau_H$
the field evolution is close to classical. At the transition between
these limiting cases, at expansion parameter $a_e$, the Hubble
parameter is
\begin{equation}
H_e\sim \lambda ^{-1/2}H_x.
\label{xe}
\end{equation}
Since the $y$-field
at $a=a_e$ is close to statistical equilibrium its characteristic
value is
$y_e\sim\lambda ^{-1/4}H_e$ (Eq.~\ref{10}). This is
related to the value of the inflaton at $a=a_e$ by
\begin{equation}
y_e/\phi _e\sim (\lambda _\phi /\lambda )^{1/2}.
\label{eqr}
\end{equation}
Kofman and Linde \cite{KL} show that, if the inflaton and
the $y$-field both have quartic potentials,
the evolution in the slow roll classical approximation is
\begin{equation}
(\lambda_\phi \phi^2)^{-1}-(\lambda y^2)^{-1} ={\rm constant}
= (\lambda_\phi \phi _x^2)^{-1}-(\lambda y_x^2)^{-1} .
\label{KoLi}
\end{equation}
Eq. (\ref{eqr}) says the constant of integration is not very
important at $a=a_e$, so a good approximation to the smaller
field values at $a_e\ll a\lower.5ex\hbox{$\; \buildrel < \over \sim \;$} a_x$ is
\begin{equation}
y(t)=(\lambda_\phi/\lambda)^{1/2}\phi(t).
\label{yphi}
\end{equation}
We consider first the field fluctuations frozen
at $a_e<a<a_x$, when the evolution of $y$ is close to classical. In
unit logarithmic interval of $x$ the contribution to the
variance of the frozen dark mass
field is $\delta y_x^2=(H/2\pi )^2$. Apart from regions where $y$
happens to be close to zero, this is a small fractional perturbation
to the field value. With Eqs.~(\ref{KoLi})
and~(\ref{yphi}) we see that the
field perturbation at the end of inflation is\footnote{To avoid confusion
we remind the reader that $\delta y_x/y_x$ is the fractional
fluctuation in the dark matter field value at the end of
inflation. The subscript $x$ in Eq.~(\ref{dyx})
has nothing to do with the
comoving length scale of the fluctuation in Eq.~(\ref{smallscalecase}).}
\begin{equation}
{\delta y_x\over y_x} = {\delta y\over y}\left( y_x\over y\right) ^2
= {H\over 2 \pi\phi _x}\left(\phi _x\over\phi\right) ^3
\left(\lambda\over\lambda _x\right) ^{1/2}.
\label{dyx}
\end{equation}
Here $H$, $y$, and $\phi$ are evaluated when field fluctuations
on the scale $x$ are frozen, at $\log (x/x_x) = (\phi /\phi _x)^2$
in the approximation of Eq.~(\ref{xoverxx}). With Eq.~(\ref{Hoft})
we get
\begin{equation}
\delta y_x/y_x\sim 0.1\lambda ^{1/2} [\log (x/x_x)]^{-1/2}.
\label{smallscalecase}
\end{equation}
Since this is nearly independent of the length scale $x$
the power spectrum varies with wave number about as $k^{-3}$,
and the mass correlation function is
\begin{equation}
\xi\sim M\lambda\log (x_e/x),
\label{xixp}
\end{equation}
where $M$ is a dimensionless constant.
At separations larger than the field coherence
length,
$R_c\sim (a_eH_e)^{-1}\exp(1/p)$ with $p\sim\lambda^{1/2}$
(Eq.~\ref{12}), Eq.~(\ref{xixp}) fails; the
values $y_e$ may be quite different
at the two points. The classical evolution Eq.~(\ref{KoLi})
says the field values at a given comoving position
$a=a_e$ and at the end of inflation
at $a=a_x$ satisfy
\begin{equation}
{1\over y_x^2} = {1\over y_e^2} + {\lambda\over H_x^2}(1-\sqrt\lambda ).
\label{ybound}
\end{equation}
If $\lambda\ll 1$ this says the field value is
$y_x\sim\pm\lambda ^{-1/2}H_x$ everywhere except near the surfaces
where $y_e$ vanishes, and the energy density is
\begin{equation}
\rho _x\sim \lambda ^{-1}H_x^4,
\label{rhox}
\end{equation}
except near the zeros of $y_e$. If the distance between zeros
were larger than the typical motion of the dark mass after
inflation the present distribution would be close to uniform
apart from surfaces of low density.
We estimate the form of the mass correlation function
at the end of inflation and for separation
$x\gg R_c$ by the argument used to obtain the
scaling of the $n$-point mass correlation functions
in the de~Sitter-invariant case in Eq.~(\ref{BB}).
Consider two windows with fixed sizes and positions
in comoving coordinates. The dark mass density contrasts
smoothed within these windows are $\bar\delta _e(1)$ and
$\bar\delta _e(2)$ at expansion parameter $a=a_e$,
when the field
fluctuations start to depart from the de~Sitter-invariant state,
and $\bar\delta _x(1)$ and
$\bar\delta _x(2)$ at the end of inflation
at $a=a_x$. If the comoving separation $a_ex_{12}$ of the windows
is large compared to $R_c$ at $a_e$
then we have, following the derivation of Eq.~(\ref{AB}),
\begin{equation}
\langle\bar\delta _x(1)\bar\delta _x2)\rangle =
T^2 \langle\bar\delta _e(1)\bar\delta _e(2)\rangle
\sim N(x_x/x_{12})^{0.6\sqrt\lambda } \sim \xi _x(x_{12}).
\label{xix}
\end{equation}
The power law coefficient follows from the de~Sitter-invariant
correlation function ~(\ref{15}). The
transfer coefficient $T$, defined as
in Eq.~(\ref{transfer_eq}), depends on $\lambda$. In general
it also depends on the window size and shape, but in the limit where
the two windows are small compared to their separation Eq~(\ref{xix})
reduces to the two-point mass correlation function.
Quite similarly, the three-point mass correlation function
satisfies
\begin{equation}
\langle\bar\delta_(1)\bar\delta_x(2)\bar\delta_x(3)\rangle =
T^3\langle\bar\delta_e(1)\bar\delta_e(2)\bar\delta_e(3)\rangle .
\end{equation}
Combined with Eq.(\ref{xix}) this implies that the
relation (\ref{18}) between the three- and two-point correlation
functions is still valid at the end of inflation.
If $\lambda$ is not much smaller than unity
the near de~Sitter evolution ends not
long before the end of inflation. In this case an estimate
of the mass correlation function at
the end of inflation from a numerical
realization of the process is easy and useful. As in Section II,
we label successive e-foldings of the comoving Hubble length by
the integer $j$, but now $j$ decreases with time, with $j=1$ at
the end of inflation. The values of the Hubble constant follow from
Eq~(\ref{simulation}),
\begin{equation}
e^{j - 1} = {x_j\over x_x} = {e^{H_j/H_x - 1}\over H_j/H_x} .
\label{Hjn}
\end{equation}
Eq.~(\ref{KoLi}) says the
values of the dark matter and inflation fields at successive
e-foldings are related by
\begin{equation}
{1\over\lambda }\left( {1\over y_j^2} - {1\over y_{j+1}^2}\right) =
{1\over\lambda _\phi }\left( {1\over \phi _j^2} - {1\over \phi
_{j+1}^2}\right).
\end{equation}
The result of multiplying this by $H_x^2$, rearranging, and adding
the quantum noise term $\pm H/(2\pi )$ at each
e-folding of $x$ is
\begin{equation}
z_j = {z _{j+1}\over [1 + G_j z_{j + 1}^2]^{1/2}} +
{\iota _j h_j\over 2\pi },
\label{Hj}
\end{equation}
where the $\iota _j$ are independent Gaussian normal numbers,
the field has been scaled to
\begin{equation}
z =y/H_x,
\end{equation}
and
\begin{equation}
G_j = {2\lambda\over 3}\left( {H_x\over H_j} - {H_x\over H_{j + 1}}\right).
\end{equation}
The $H_j$ come from the numerical solution of
Eq.~(\ref{Hjn}). The two-point function at separation
$x/x_x=e^j$ is the mean of products of field values computed
starting from a common value $j$ time steps before
the end of inflation.
Figure~3 compares the mass correlation
functions at the end
of inflation for constant Hubble parameter and for the rolling
case in Eq.~(\ref{Hoft}), for $\lambda = 0.1$.
One sees that at relatively small separations $\xi (x)$ is
considerably flatter in the rolling $H$ case, as
expected from Eq~(\ref{xixp}). The
values and rates of change of the correlation functions
are roughly similar at separations $x\sim 10^{15}x_x$.
This suggests that for $\lambda =0.1$
the constant $N$ in Eq.~(\ref{xix}) is not greatly
different from unity.
At smaller $\lambda$ a numerical realization of the
two-point function is difficult because
the relaxation time is long, but it is easy to get the
one-point distribution of $y$ at the end of inflation.
At $\lambda =0.001$ the distribution is strongly
peaked at $y\sim\pm\lambda ^{-1/2}H_x$,
as expected (Eq.~\ref{rhox}), but there is
significant scatter from the quantum fluctuations
appearing at $a\lower.5ex\hbox{$\; \buildrel < \over \sim \;$} a_e$. For $\lambda =0.1$ the distribution
of field values $y$ is peaked at zero, and not greatly different
from the stationary de~Sitter case.
In the next section we use the estimates of the
field mass fluctuations in Eqs.~(\ref{smallscalecase})
and~(\ref{xix}) with the measured CBR anisotropy to
find bounds on the parameter $\lambda$. We will argue that the
considerable uncertainty in $N$ translates to a relatively small
uncertainty in the bound on $\lambda$.
\section{The Dark Matter at the Present Epoch}
We consider three constraints: the present mean mass
density in dark matter, the small-scale cutoff in the
gravitational growth of clustering of the $y$-matter, and the
large-scale anisotropy of the thermal background radiation.
Because most of our estimates are quite approximate we
ignore most numerical factors that are of order unity.
\subsection{The Mean Mass Density}
With the change from proper world time $t$ to
conformal time $x^0=\int dt/a$, and the definition
$\tilde y=ay$, the action
for the $y$-field is
\begin{equation}
S = \int d^4x\left( {1\over 2}\tilde y_{,i}\tilde y^{,i}
-{1\over 4}\lambda\tilde y^4 - {1\over 12}a^2\tilde y^2R\right) ,
\label{19a}
\end{equation}
where the index is raised by the Minkowski metric tensor and we ignore
the quadratic part of the potential.
The Ricci tensor $R$ is on the order of the square of the Hubble
parameter $\dot a/a$ when the universe is matter-dominated, and it is
well below that when radiation-dominated. When the Ricci term
may be neglected a solution for $\tilde y$ in Minkowski
spacetime is a solution for $ay$ expressed as a function of
comoving coordinates and conformal time. The $y$-field energy
density depends on the proper time derivative
$\dot y = (\partial\tilde y/\partial x^0 - \tilde y\dot a)/a^2$.
If the field oscillates rapidly on the scale of the Hubble time,
either because the potential is driving rapid oscillations or the field
varies with position on scales small compared to the Hubble length,
then $\dot y=a^{-2}\partial\tilde y/\partial x^0$ to good accuracy, and
the proper field energy density is well approximated
as the energy density of
$\tilde y$ in Minkowski coordinates divided by $a^4$. Since the
former conserves energy the mean
energy density in the $y$-field varies as
\begin{equation}
\rho _y\propto a^{-4}.
\label{22}
\end{equation}
The characteristic amplitude of the field oscillations thus
scales as
\begin{equation}
y\propto 1/a(t).
\label{23}
\end{equation}
Ford\cite{Ford} obtained these results
for a quartic potential when $y$ is a function only of world
time. It is an easy exercise to obtain Eqs~(\ref{22})
and~(\ref{23}) by extending Ford's method to
the case where the scales of length and time
variations of $y$ both are small compared to $H^{-1}$.
The mean field value at the end of inflation is, from
Eqs.~(\ref{Hoft}) and~(\ref{yphi}),
\begin{equation}
y_x\sim \lambda ^{-1/2}H_x,
\label{y_xvalue}
\end{equation}
At this field value
the characteristic field oscillation frequency
is $\omega\sim\lambda ^{1/2}y_x\sim H_x$.
This means the field starts oscillating at about the time
inflation ends, so the field amplitude after inflation decays as
$y\sim y_xa_x/a$ until the quadratic part of the potential
becomes comparable to the quartic part, at
$y\sim y_\mu=\mu\lambda ^{-1/2}$. The expansion factor at which
this happens is
\begin{equation}
a_\mu/a_x\sim H_x/\mu .
\label{23a}
\end{equation}
The present mean mass density in the $y$-field, which
we are assuming is comparable to the total, is
\begin{equation}
\rho _o\sim z_{\rm eq}T_o^4\sim \lambda ^{-1}
H_x^4(a_x/a_\mu )^4(a_\mu /a_o)^3.
\label{24}
\end{equation}
The mass density at the end of inflation is
$\lambda y_x^4/4\sim\lambda ^{-1}H_x^4$ (Eq~\ref{y_xvalue}).
The present CBR temperature is $T_o$, and
the redshift at equality of mass densities in
matter and radiation is
$z_{\rm eq}\simeq 3500$ (for
$\Omega _m = 0.3$ and $h = 0.7$).
In the conventional model for reheating
it will be sufficient to consider the case
where the expansion is radiation-dominated at
temperature $T_x\sim (m_{\rm pl} H)^{1/2}$
at the end of inflation. Then the expansion
factor from the end of inflation to
the present is $a_o/a_x\sim T_x/T_o$. With Eqs.~(\ref{23a})
and~(\ref{24}) we get
\begin{equation}
\mu\sim \lambda z_{\rm eq}T_om_{\rm pl} ^3/T_x^3\sim 10^5\lambda
T_{14}^{-3}\hbox{ GeV},
\label{25}
\end{equation}
for temperature $T_x=10^{14}T_{14}$~GeV at the end of inflation.
The redshift at which the dark matter starts acting as a massive
field is
\begin{equation}
a_o/a_\mu\sim\lambda z_{\rm eq}(m_{\rm pl} /T_x)^4\sim 10^{22}\lambda T_{14}^{-4}.
\label{26}
\end{equation}
If $\lambda\gg 10^{-14}$ and $T_{14}\sim 1$
the length scales of interest for extragalactic
astronomy appear at the Hubble
length at redshifts well below this value of $a_o/a_\mu$,
and the dynamical behavior of the $y$-matter is not
significantly different from the usual cold dark matter.
In the quintessential picture the mass densities at the end of
inflation in gravitationally produced interacting matter and
noninteracting matter are
\begin{equation}
\rho_x({\rm matter}) = CH_x^4,\qquad
\rho _x({\rm dm})\sim\lambda ^{-1}H_x^4.
\label{mass_densities}
\end{equation}
The constant $C$ depends on the matter interactions, as discussed
in Section VI-A. Consistency with the standard model
for the origin of the light elements requires that the ratio
of these mass densities, $f\sim\lambda C$,
be greater than about $f\sim 15$ at the time of light element
production. If this condition is violated the model is excluded.
If the model is viable we can normalize the parameter constraints
to the value of $f$. The redshift at which the $y$-field starts acting
like a nonrelativistic massive field is
\begin{equation}
z_\mu = a_o/a_\mu\sim f z_{\rm eq}.
\label{27}
\end{equation}
The radiation temperature at the end of inflation is
$T_x\sim C^{1/4}H_x$, so the expansion factor to the present is
$a_o/a_x\sim C^{1/4}H_x/T_o$.
With Eq. (\ref{23a}) we have
\begin{equation}
\mu\sim \lambda C^{3/4}z_{\rm eq}T_o \sim 1\lambda C^{3/4}\hbox{ eV}.
\label{28}
\end{equation}
\subsection{The Effective Jeans Length}
In quintessential inflation $y$ starts acting like a massive
field at relatively low redshift (Eq.~\ref{27}), and we have to check
the effect on the gravitational growth of clustering of the
dark matter. The $y$-field fluctuations
that appear at the Hubble length well after inflation
have wavelengths that are much
larger than the period of oscillation of $y$, and the
gradient energy density $(\nabla y)^2/2$ thus is strongly
subdominant to the kinetic and potential energy densities. We
assume dynamical evolution leaves the gradient energy subdominant.
We proceed by deriving a virial theorem valid when the field
oscillation frequency is much larger than the Hubble parameter,
following Ford\cite{Ford}, and use it to estimate the
effective pressure from the trace of the stress-energy tensor.
Neglecting the cosmological expansion, the field equation is
\begin{equation}
\ddot y - \nabla ^2y + \lambda y^3 + \mu ^2y = 0.
\label{31}
\end{equation}
The result of multiplying this equation by $y$, averaging over space,
and averaging over a time interval that is much longer than the field
oscillation time and much shorter than the cosmological expansion time
is
\begin{equation}
\langle\dot y^2\rangle = \langle (\nabla y)^2\rangle + \lambda\langle y^4\rangle
+\mu ^2\langle y^2\rangle .
\label{32}
\end{equation}
With this relation the mean energy density is
\begin{equation}
\langle\rho\rangle = \langle (\nabla y)^2\rangle + 3 \lambda\langle y^4\rangle /4
+ \mu ^2\langle y^2\rangle ,
\label{33}
\end{equation}
and the mean of the trace of the stress-energy tensor is
\begin{equation}
\langle T\rangle = \langle\rho\rangle - 3\langle p\rangle = \mu ^2\langle y^2\rangle ,
\label{34}
\end{equation}
where $p$ is the effective pressure. If the gradient energy is
subdominant we have
\begin{equation}
{\langle p\rangle\over\langle\rho\rangle } = {1\over 3}{\lambda\langle y^4\rangle\over
\lambda\langle y^4\rangle + 4\mu ^2\langle y^2\rangle /3}.
\label{35}
\end{equation}
Properties of $y(t)$ considered as a single oscillator with
the equation of motion
$\ddot y=-\lambda y^3$ are discussed by Greene, Kofman, Linde,
and Starobinsky \cite{Greene}. The energy equation for the
oscillator is
\begin{equation}
\dot y^2/2 + \lambda y^4/4 = \lambda y_o^4/4,
\label{36}
\end{equation}
where the amplitude is $y_o$, and we can use this
to compute time averages of moments of $y$. In particular,
\begin{equation}
\langle y^2\rangle ^2/\langle y^4\rangle = 48[\Gamma (3/4)/\Gamma (1/4)]^4 \simeq 0.63.
\label{37}
\end{equation}
When the amplitude is small enough that the
quadratic part of the potential dominates,
\begin{equation}
\langle y^2\rangle ^2/\langle y^4\rangle = 2/3.
\label{38}
\end{equation}
In short, we can take it that $\langle y^4\rangle\sim \langle y^2\rangle ^2$,
and we see from Eq.~(\ref{35}) that
$\langle p\rangle /\langle\rho\rangle\simeq 1/3$ until
$\langle y^2\rangle\sim\mu ^2/\lambda$, at expansion parameter $a\sim a_\mu$.
In quintessential inflation this is at redshift
$z_\mu\sim fz_{\rm eq}$ (Eq.~\ref{27}). At $z<z_\mu$ the
ratio varies as $\langle p\rangle /\langle\rho\rangle\propto a^{-3}$, because
$\langle y^2\rangle\propto a^{-3}$ for a massive nonrelativistic field.
Thus the effective velocity of sound is
$c_s\propto a^{-3/2}\propto t^{-3/4}$
at $z_\mu >a>z_{\rm eq}$ and $c_s\propto t^{-1}$
at $z<z_{\rm eq}$. The physical Jeans length
$\sim c_st$ increases as $t^{1/4}$ at $z_\mu\lower.5ex\hbox{$\; \buildrel > \over \sim \;$} z\lower.5ex\hbox{$\; \buildrel > \over \sim \;$} z_{\rm eq}$
and then approaches a constant. The comoving Jeans length
$c_st/a$ is maximum at $z_\mu$, and the maximum comoving Jeans
length referred to the present epoch is
\begin{equation}
L _{\rm J}\sim t_\mu z_\mu /\sqrt{3}\sim 6/(f\Omega _m h^2)
\hbox{ Mpc}.
\label{39a}
\end{equation}
The world time at redshift $z_\mu$ is $t_\mu$, the present matter density
parameter is $\Omega _m$, and the present Hubble parameter is
$h$. On comoving scales larger than
$L_{\rm J}$ it is a good approximation to
neglect the effect of the field stress on
the gravitational evolution of the $y$-field mass distribution.
On smaller scales
the dynamical behavior of the field requires a more
detailed analysis than is attempted here.
The effective Jeans length in Eq.~(\ref{39a}) is smaller than that
of a family of neutrinos with mass $\sim 30$~eV,
and may be comparable to the mean distance between large
galaxies.
\subsection{The Large-Scale Thermal Background Anisotropy}
The parameter $\lambda$ is bounded by the effect of
isocurvature fluctuations in the $y$-field at the end of
inflation on the angular distribution of the thermal
background radiation (the CBR). For definiteness we present
numerical results for quintessential inflation.
This analysis of the temperature fluctuations
on large angular scales uses the simplification
that the radiation pressure
gradient force has little effect on the mass distribution
on the scale of the present
Hubble length, so we can imagine
that well-separated regions we see at the Hubble length evolve
as separate homogeneous cosmological
models. In one of these homogeneous models the field value and
radiation temperature at the end of inflation are $y_x$ and
$T_x$. By repeating the computation in Section IV-A one finds
that the present dark mass density varies with these parameters and
the present temperature, $T_o$, as
\begin{equation}
\rho _o \sim\lambda ^{1/2}C^{-3/4}\mu (y_xT_o/T_x)^3.
\label{39}
\end{equation}
Under the isocurvature initial conditions (and assuming the
$y$-mass density at high redshift
is subdominant, as required for light element
production), the temperature $T_x$ is nearly
homogeneous, the same in all
the model universes that represent the evolution of different
regions. Also, the present dominant mass density,
$\rho _o$, is close to homogeneous on the scale of the present Hubble
length, because pressure gradient forces cannot
generate large-scale curvature fluctuations. Thus the present
large-scale CBR anisotropy is
\begin{equation}
{\delta T_o\over T_o}\simeq - {\delta y_x\over y_x} \simeq
-{1\over 4}{\delta\rho _x\over\rho _x}.
\label{40}
\end{equation}
This relation also follows in the conventional model for reheating.
We need expansion factors during and after inflation.
In quintessential
inflation the ratio of the comoving Hubble length now,
$x_o = (H_oa_o)^{-1}$, and the Hubble length $x_x$ at the end
of inflation is
\begin{equation}
{x_o\over x_x}\sim {H_xa_x\over H_oa_o}\sim C^{-1/4}{T_o\over H_o}.
\label{46}
\end{equation}
The ratio of the present temperature and Hubble parameter is
\begin{equation}
T_o/H_o\sim e^{67}.
\label{TooverHo}
\end{equation}
When field fluctuations during inflation
are being frozen at the comoving
scale $x_o$ of the present Hubble length
the expansion and Hubble parameters
$a_p$ and $H_p$ satisfy
\begin{equation}
H_pa_p\sim H_oa_o\sim C^{1/4}H_xa_xH_o/T_o.
\end{equation}
These relations with Eq.~(\ref{axovera}) say
the expansion factor from $a_p$ to the end of
inflation satisfies
\begin{equation}
a_x/a_p\sim C^{-1/4}(T_o/H_o)\log a_x/a_p.
\end{equation}
If $C$ is on the order of unity the numerical solution is
\begin{equation}
H_p/H_x = \log a_x/a_p \simeq 72.
\label{HpoverHx}
\end{equation}
We have to consider two cases, where the dark matter field
departs from near statistical equilibrium
and commences classical slow roll before or after
$ax_o$ appears at the Hubble length during inflation.
The Hubble parameter
at the breaking of equilibrium is
$H_e\sim\lambda ^{-1/2}H_x$ (Eq.~\ref{xe}), so a critical
parameter value is
\begin{equation}
\lambda _p =(H_x/H_p)^2\sim 1\times 10^{-4},
\label{lambdap}
\end{equation}
from Eq~(\ref{HpoverHx}).
If $\lambda\ll\lambda _p$ then we observe
in the CBR anisotropy the
effect of fluctuations frozen while $y$ was behaving as a
classical field that is slightly perturbed by quantum
fluctuations. In this case
Eqs.~(\ref{smallscalecase}), (\ref{40}) to~(\ref{TooverHo})
and~(\ref{HpoverHx}),
with the measurement $\delta T_o/T_o\simeq 1\times 10^{-5}$,
imply
\begin{equation}
\lambda\lower.5ex\hbox{$\; \buildrel < \over \sim \;$} 1\times 10^{-6}.
\label{smalllambdalimit}
\end{equation}
This assumes $C$ is on the order of unity in Eq.~(\ref{46}),
and unless $C$ is very large its value does not significantly
affect the bound on $\lambda$. The condition $C\sim f/\lambda$
from Eq.~(\ref{mass_densities}) with $f\sim 10$
requires $C\lower.5ex\hbox{$\; \buildrel > \over \sim \;$} 10^7$. This limit is close enough
to unity for the purpose of the numerical estimate
in Eq.~(\ref{HpoverHx}).
The bound in Eq.~(\ref{smalllambdalimit}) differs from the
analysis of Felder, Kofman and Linde \cite{fkl}
by the factor in Eq.~(\ref{dyx}) that takes account of the
classical decay of $y$ from $a_p$ to the end of inflation.
In the other limiting case,
$\lambda\gg\lambda _p$, it is convenient to use the expansion
of the observed angular distribution of the background
temperature in spherical harmonics, with coefficients
\begin{equation}
a_l^m= \int d\Omega\, Y_l^m \delta T_o/T_o.
\label{41}
\end{equation}
With Eq~(\ref{40}), and using the addition theorem,
$\sum _mY_l^m(1)Y_l^{-m}(2)=(2l+1)P_l(\cos\theta _{12})/4\pi$,
we can write the mean square value of a multipole moment as
\begin{equation}
\left(\delta T_l\over T\right) ^2 = {l(2l+1)\over 4\pi }\langle |a_l^m|^2\rangle =
{l(2l+1)\over 4\pi }\int d\Omega P_l(\cos\theta )\xi (x)/16.
\label{42}
\end{equation}
In this normalization, $(\delta T_l/T)^2$ is the contribution to
the variance of the sky temperature per logarithmic interval of
the spherical harmonic index $l$.
The argument of the dark mass correlation function at the end
of inflation is $x = 2x_o\sin\theta /2$, where the present
angular size distance $x_o$ back to high redshift is
\begin{equation}
x_o\simeq 3.2(H_oa_o)^{-1}.
\label{angularsizedistance}
\end{equation}
Here $\Omega _m=0.3$, we assume zero space curvature,
and the present Hubble and expansion
parameters are $H_o$ and $a_o$.
For the dark mass correlation
function in Eq.~(\ref{xix}) the large-scale anisotropy spectrum is
\begin{equation}
\left(\delta T_l\over T\right) ^2 = N {l(2l+1)\over 32}
\left( x_x\over x_o\right) ^{0.6\sqrt{\lambda }}\int _{-1}^1 d\mu
P_l(\mu ) (2 - 2\mu )^{-0.3\sqrt{\lambda }}\lower.5ex\hbox{$\; \buildrel < \over \sim \;$} 1\times 10^{-10}.
\label{spect}
\end{equation}
The observational bound applies at $l\lower.5ex\hbox{$\; \buildrel < \over \sim \;$} 30$.
We have written $x_x/x=(x_x/x_o)(x_o/x)$; the second factor
produces the last factor in the integrand. The solution to this
equation at $\lambda\ll 1$ is not relevant because the limit in
Eq.~(\ref{smalllambdalimit}) applies. In the solution at
$\lambda$ close to unity the integral is of order unity and the
value of the integral and the factor $N$ in the primeval mass
correlation function do not much matter because the sensitive
function is $(x_x/x_o)^{0.6\lambda }$. We have
\begin{equation}
\lambda \lower.5ex\hbox{$\; \buildrel > \over \sim \;$} 0.3.
\label{largelambdalimit}
\end{equation}
In the conventional reheating picture $\log a_x/a_p$ is about ten
percent smaller. To the accuracy of our estimates the constraints
on $\lambda$ are the same.
\section{DISCUSSION}
We comment first on issues of consistency and reasonableness
of the noninteracting dark matter
picture within conventional and quintessential endings
of inflation, and then take note of some observational challenges.
\subsection{Theoretical Issues}
For the purpose of exploring a simple example of dark matter
that interacts only with itself and gravity
we have considered a quartic plus quadratic
self-interaction potential.
This functional form has the arguably attractive feature that one can
choose constant coefficients that imply an interesting present mean
dark mass density and CBR anisotropy.
We are not competent to say whether other
models for a self-interaction
potential would be more plausible from the point of view of
fundamental physics or would produce dark matter
candidates with more interesting properties.
In the quartic plus quadratic potential the isotropy of the CBR
requires that the dimensionless parameter $\lambda$ be in the
narrow range
\begin{equation}
0.3\lower.5ex\hbox{$\; \buildrel < \over \sim \;$}\lambda\lower.5ex\hbox{$\; \buildrel < \over \sim \;$} 1,
\label{52}
\end{equation}
or else be quite small,
\begin{equation}
10^{-6}\lower.5ex\hbox{$\; \buildrel > \over \sim \;$}\lambda\gg\lambda _\phi\lower.5ex\hbox{$\; \buildrel < \over \sim \;$} 10^{-14}.
\label{secondrange}
\end{equation}
The last bound is the condition on the adiabatic
density fluctuations produced by the inflaton.
The upper limit in Eq.~(\ref{52}) is the
condition that the potential be safe
from significant renormalization (Eq.~\ref{2'}).
We know of no reason to
think either range of values of $\lambda$
is particularly attractive within fundamental physics, though
it is to be hoped that input from this direction eventually will
be a factor in the completion of a satisfactory theory of
structure formation.
In a recent paper Felder, Kofman and Linde\cite{fkl}
(hereafter FKL) raise a number of issues relevant
to our analysis. They note that scalar fields
whose particles are produced gravitationally at the end of inflation
could end up dominating the energy density, thus assuming the role of
the inflaton. This could indeed happen if the self-couplings of
some of these fields in quartic potential models were smaller
than that of the inflaton. One sees from Eqs.~(\ref{eqr})
and~(\ref{yphi}) that the ratio of energy densities in the inflaton and
a field with self-coupling $\lambda _j$ in the quartic model is
\begin{equation}
\rho _j/\rho _\phi = \lambda _\phi/\lambda _j.
\label{condition}
\end{equation}
Thus consistency within this model requires
$\lambda_j \gg \lambda_\phi$.
Since the interacting scalar
fields of the quintessential model \cite{quint} are usual spin-0
particles, like the Higgses that interact with gauge fields,
their couplings are not expected to be particularly small.
The inflaton coupling, on the other hand, is required to be small
[Eq.(\ref{secondrange})], and there does not seem to be a problem in
meeting the conditions on $\lambda_j$.
FKL note that the $\lambda _j$ are also constrained by the
condition that the isocurvature
fluctuations produced by the scalar fields
not violate the CBR isotropy.
The point is valid when some of the fields represent
stable non-interacting massive particles. Indeed, we find
that the range $10^{-6}\ll\lambda _j\lower.5ex\hbox{$\; \buildrel < \over \sim \;$} 0.3$ is
excluded for noninteracting dark matter.\footnote{Our larger
lower bound on $\lambda _j$ seems to be mainly due to the correction
for the decay of field fluctuations in Eq~(\ref{dyx}). Our
larger allowed range applies
when the equilibrium between quantum excitations and classical
decay of the dark matter field persists to close to the end of
inflation, a case FKL do not consider.}
Interacting scalar fields, such as the Higgses, also are
produced with inhomogeneous distributions at the end of
inflation. These particles interact by the usual decay,
annihilation and production processes that produce local thermal
equilibrium. In the absence of particle number conservation laws,
this local thermal equilibrium is fully
described by a single function of position --- the
temperature or the mass density. Apart from the adiabatic
perturbations from the inflaton
the model predicts negligible spacetime
curvature fluctuations at the end of inflation. This means
the matter density fluctuations correspond to fluctuations
in the local starting
times of cosmological expansion with universal values of the
cosmological parameters. If all matter interacts and relaxes to
local thermal equilibrium the result at the present time is an
unacceptably homogeneous universe. If a dark matter component
does not interact with the rest of the matter, the possibility
considered in this paper, fluctuations in composition, which
is to say isocurvature fluctuations, remain. There may also be
adiabatic curvature fluctuations from the inflaton, of course.
Either or both could act as seeds for structure formation.
FKL also point out that the moduli problem is more severe than in
the conventional picture for reheating.
Moduli are light scalar fields with only gravitational-strength
coupling to ordinary matter. They necessarily arise in
supergravity models and superstring theories. The
curvature correction generally makes the effective potential of
a modulus $\chi$ during inflation different from at
late times, the minima of the two potentials tending to be
displaced by $\Delta\chi\simm_{\rm pl}$. Thus, after
inflation one has a nearly homogeneous field $\chi$ with $\rho_\chi
\sim m_\chi^2(\Delta\chi)^2\sim m_\chi^2m_{\rm pl} ^2$. The mass $m_\chi$
is determined by the supersymmetry breaking scale $\eta_{SUSY}$,
\begin{equation}
m_\chi\sim\eta_{SUSY}^2/m_{\rm pl}.
\end{equation}
In models with gravity-mediated supersymmetry breaking,
$\eta_{SUSY}\sim 10^{11}~$~GeV and $m_\chi\sim 10^3$~GeV.
The main problem with the moduli is
that they may decay very late.
With $m_\chi\sim 10^3$~GeV, the lifetime is
\begin{equation}
\tau\sim m_{\rm pl}^2/m_\chi^3\sim 10^5 ~s,
\label{lifetime}
\end{equation}
so one runs into problems with nucleosynthesis and with
photodissociation and photoproduction of light elements by the decay
products of $\chi$\cite{Ellis}.
There are two commonly discussed possible solutions
to the moduli problem. First,
a short period of secondary inflation could
dilute the moduli\cite{Randall}. This naturally
occurs in ``thermal'' inflation\cite{Lyth}, but
seems unlikely to be effective in quintessential inflation,
where the relative density of the moduli is much
higher than in the conventional picture, as pointed out by
FKL. Second, the minima of $V(\chi)$ during and after
inflation may coincide due to some symmetry\cite{Dvali,Randall2,Damour}.
In this case moduli are produced only gravitationally, like other
light scalars, and have density $\rho_\chi\sim H^4$ at the end of
inflation. The density of matter in standard inflation is $\rho_m\sim
m_{\rm pl}^2 H^2$, and $\rho_\chi/\rho_m\sim H^2/m_{\rm pl}^2$ is small enough to solve
the
moduli problem. However, in quintessential inflation $\rho_m\sim H^4$
and $\rho_\chi/\rho_m\sim 1$, which is clearly too high.
There are other ways out, however. During inflation, moduli typically
acquire masses \cite{Dvali,Dine} $m_\chi\sim \beta H$ with $\beta\gtrsim 1$,
and it is conceivable that $\beta\gg 1$ \cite{Dvali,Linmod}. The
gravitational production of moduli would then be exponentially
suppressed. Moduli can also develop non-perturbative potentials,
independent of supersymmetry breaking \cite{DK}. Their mass can then
be large enough that the lifetime in
Eq.~(\ref{lifetime}) is $\tau \ll 1 s$. Assuming coincident minima of
$V(\chi)$ during and after inflation, this would eliminate the
problem with light elements. If the moduli decay prior to
baryogenesis, then the relaxation to thermal
equilibrium eliminates the isocurvature perturbation
associated with the inhomogeneous primeval local ratio
of entropy density to the number density of modulus
quanta. Otherwise, the isocurvature perturbation survives in the
form of an inhomogeneous ratio of the baryon and photon number
densities and can later play a role in structure formation.
Another possibility is that the supersymmetry breaking scale could be
much smaller than $10^{11}$~GeV; in some models it can be as low as
the electroweak scale \cite{ADD}.
The moduli would then be very light, their lifetime
would be much greater than the present age of the
universe, and problems with nucleosynthesis would not arise.
(Here we still assume coincident minima and also that the moduli
develop masses $\sim H$ during inflation).
There is a related issue for gravitinos. The lower energy
density at the end of inflation in the quintessential
picture reduces thermal gravitino production
but increases the problem of gravitational production, as
FKL note. Apart from potential energy terms the gravitino obeys
a conformally invariant field equation \cite{Grishchuk}. If during
inflation
the gravitino had a substantial effective mass,
the gravitino energy density from gravitational particle
production could be
unacceptable in quintessential inflation. We are not aware that
the effective mass has to be this large, however.
The moduli and gravitino quanta are close approximations
to the proposed noninteracting dark matter, but with potentially
unacceptable production of mass density. This is a challenge,
but within the not inconsiderable uncertainties of a
supersymmetric theory that has not yet been fully
specified the challenge does not seem serious enough to
discourage further exploration of the picture.
We note finally that the density of matter produced in
quintessential inflation can be substantially higher than our original
estimate in Ref.\cite{quint}. The quanta of
spin-0 fields $\chi_j$ produced due to the change in the expansion law
at the end of inflation have wavelength $\sim H_x^{-1}$ and energy
density $\rho_j^{(s)}\sim H_x^4$. In addition to these
short-wavelength quanta, there is also a nearly homogeneous component
of $\chi_j$. The classical solution (\ref{yphi}) always applies at
the end of inflation, and the energy of the homogeneous component
follows as in Eq.~(\ref{mass_densities}),
\begin{equation}
\rho_j^{(h)}\sim(\lambda_\phi^2/\lambda_j)m_{\rm pl}^4\sim\lambda_j^{-1} H_x^4,
\end{equation}
where $H_x\sim \lambda_\phi^{1/2}m_{\rm pl}$ is the expansion rate at the end
of inflation. We see that $\rho_j^{(h)}\gg\rho_j^{(s)}$ for small
$\lambda_j$. For ordinary (interacting) matter fields, the
homogeneous component decays and thermalizes, just like the
short-wavelength quanta. We note also that this
enhancement of the energy density in the homogeneous component
disappears if the fields $\chi_j$ develop masses $m_j\sim H$ during
inflation. In this case the original estimate of \cite{quint},
$\rho_j\sim H_x^4$, is still valid, for $\lambda_j < 1$,
because the quartic term is small compared to the quadratic part
of the potential.
\subsection{Observational Issues}
Previous examples of noninteracting dark
matter \cite{LindeMuk}, \cite{Peeb} postulate a quadratic
self-interaction potential with a time-variable mass.
These are CDM models, where the CDM is defined as matter
that has had negligible pressure since appearance at the
Hubble length of all length scales of astrophysical interest.
The two-field CDM model in \cite{Peeb} could be adjusted to fit the
spectrum $\delta T_l/T$ of angular fluctuations of the thermal
background, to the accuracy of the measurements then available,
but at $l\sim 50$ the model power is a factor of two above the
new Python\cite{Python} result. Bridging this
gap would require even greater contrivance. This
isocurvature CDM model thus does not seem
to be viable, but it does offer a useful reference for
a first assessment of observational tests of the quartic plus
quadratic self-interaction model.
In the isocurvature CDM model the primeval distribution of
the radiation is much smoother than that of the CDM because the
mean mass density in the CDM is much smaller. With
the quartic potential considered here the dark matter
approximates a fluid with the equation of state of the
radiation until the redshift approaches $z_{\rm eq}$
(Eq.~\ref{27}). This means
a given primeval isocurvature fluctuation
$\delta\rho _y/\rho _y$ in the dark matter
distribution produces a much larger primeval perturbation to the
distribution of the radiation than for CDM.
The effect on the fluctuation
spectrum of the CBR on angular
scales below the limit of application of Eq.~(\ref{spect})
remains to be computed.
In the isocurvature CDM model the primeval mass distribution is
$\rho\propto y ({\bf x})^2$, with $y$ a random Gaussian
process, and the non-Gaussian mass density
fluctuations in this model may violate the measured
skewness and kurtosis of the large-scale fluctuations of
galaxy counts\cite{Gaz}. The numerical realizations in
Fig.~2 indicate that in the quartic potential
model the skewness of the primeval mass distribution
is a factor of about 3 smaller than in the case
$\rho\propto y ^2$.
Again, the effect on the observational
constraint may be worth investigating.
In quintessential inflation the
dynamical behavior of the $y$-field would be an important factor
in the formation of the first generations of structure
(Eq.~\ref{39a}). Khlebinkov and Tkachev\cite{Tkachev} present
numerical simulations of the three-dimensional behavior of a
scalar field with quartic self-interaction potential, in
connection with the dynamics of reheating in conventional
inflation. They find a marked tendency for the growth of
field fluctuations on ever smaller scales. The qualitative effect
is easily understood: where the field value $|y|$ is larger than
average the field oscillation frequency is larger, and the
frequency differences at different positions produce growing field
gradients. Our preliminary numerical experiments in one space
dimension plus time suggest the growing field gradients produce
energy fluxes that redistribute energy and the space distribution
of field oscillation frequencies. This tends to unwind the field
gradients through substantial parts of space, while leaving
localized regions with relatively large field gradient energy
density.
The effect is curious enough that further study would be at
least intellectually interesting, and it is
conceivable that it adds an interesting
complication to the observed growth of structure on the scale
of galaxies.
Finally, we return to our opening remark, that it may be
useful to complement the growing observational basis for a theory
of structure formation with surveys of the
options available within conventional physics.
It is natural to consider simplest possibilities first,
but prudent to bear in mind the possibility of complications. An
example is the evidence that the density parameter in
matter capable of clustering on scales much smaller
than the Hubble length is significantly below unity\cite{Omega}.
If this were valid it would mean that the large-scale nature of
the universe is not as simple as we could imagine within conventional
physical ideas, a result that could hardly be described as
surprising in light of the ample opportunities
for complexity in the physical universe. We arrived at a simple
model for structure formation, adiabatic CDM, fifteen years
ago, after abandoning a few even simpler
ideas. In the present paper we have considered a class of dark matter
candidates whose properties could be considerably more
complicated than cold dark matter, and the observational
predictions much harder to work out. On the other hand,
the picture does not seem unnatural: possibly some of the fields
gravitationally produced by
inflation are unable to thermalize with ordinary matter;
perhaps such fields end up with an observationally
interesting dark mass density; perhaps their irregular
primeval distributions are a significant factor in seeding the
gravitational growth of structure.
\section{Acknowledgments}
We thank Lev Kofman for an advance copy of his recent paper with
Felder and Linde \cite{fkl}; it helped improve our paper.
We have also benefitted from
discussions with Gia Dvali, Massimo Giovannini, Andrei
Gruzinov, Lev Kofman, Andrei Linde, Alison Peebles, David Spergel,
and Paul Steinhardt. This research was
supported in part at the Institute for Advanced Study
by the Alfred P. Sloan Foundation and at Tufts University by the
National Science Foundation.
\section{Appendix A. Calculation of two- and three-point functions}
To calculate the two- and three-point correlation functions in de
Sitter space, it will be convenient to introduce the conditional
probability $\Pi(y,t|y_0,t_0)$ for the field to have value $y$ at time
$t$, given that it had value $y_0$ at time $t_0$ at the same co-moving
position. This probability satisfies the Fokker-Planck equation
(\ref{FP}) with the initial condition
\begin{equation}
\Pi(y,t_0|y_0,t_0)=\delta(y-y_0),
\end{equation}
and can be expressed as \cite{Risken}
\begin{equation}
\Pi(y,t|y_0,t_0)=e^{-v(y)+v(y_0)}\sum_{k=0}^\infty
\Phi_k(y)\Phi_k(y_0) e^{-\Lambda_k(t-t_0)}.
\label{Pi}
\end{equation}
Here,
\begin{equation}
v(y)\equiv {4\pi^2\over{3H^4}}V(y),
\end{equation}
and $\Phi_k(y)$ form a complete orthonormal set of eigenfunctions of
the equation
\begin{equation}
\left[-{\partial^2\over{\partial y^2}}+v'(y)^2-v''(y)\right]\Phi_k(y) =
{8\pi^2\Lambda_k \over{H^3}}\Phi_k(y).
\end{equation}
The eigenvalues $\Lambda_k$ are non-negative, with the smallest
eigenvalue $\Lambda_0=0$ corresponding to
\begin{equation}
\Phi_0(y)=N^{-1/2}e^{-v(y)} ,
\label{Phi0}
\end{equation}
with
\begin{equation}
N=\int_{-\infty}^\infty e^{-2v(y)}dy.
\end{equation}
Note that $\Phi_0(y)$ is related to the ``equilibrium'' distribution
function (\ref{P0}),
\begin{equation}
P_0(y)=\Phi_0^2(y).
\label{equil}
\end{equation}
we shall assume that the eigenfunctions are ordered so that
$0<\Lambda_1 < \Lambda_2 < ...~$. For a purely quartic potential
\cite{SY},
\begin{equation}
\Lambda_1 = 0.0889\lambda^{1/2}H, ~~~~~~ \Lambda_2 = 0.289\lambda^{1/2}H.
\end{equation}
Let us now consider two co-moving positions ${\bf x_1}$ and ${\bf
x_2}$ separated by a distance $ax=a|{\bf x_1-x_2}|\gg H^{-1}$ at some
moment $t=0$. These positions were within each other's horizons prior
to the time
\begin{equation}
t_a=-H^{-1}\ln (Hax).
\label{ta}
\end{equation}
For $t<t_a$, the values of $y$ at ${\bf x_1}$ and ${\bf x_2}$ are
essentially the same. Hence, the probability for $y$ to take a value
$y_1$ at ${\bf x_1}$ and a value $y_2$ at ${\bf x_2}$ at time $t=0$
can be expressed as \cite{SY}
\begin{equation}
P_2[y_1({\bf x_1}),y_2({\bf x_2})]=\int\Pi(y_1,0|y_a,t_a)
\Pi(y_2,0|y_a,t_a) P_0(y_a)dy_a,
\label{p2}
\end{equation}
where $P_0(y)$ is the equilibrium distribution (\ref{equil}). The
equal-time two-point correlation function for $y^n$ is given by
\begin{eqnarray}
\langle y^n({\bf x_1},0)y^n({\bf x_2},0)\rangle &\equiv & \langle y_1^n y_2^n\rangle
=\int_{-\infty}^\infty dy_1 dy_2 y_1^ny_2^n P_2[y_1({\bf
x_1}),y_2({\bf x_2})]\nonumber\\
&=& N^{-1}\sum_{k=0}^\infty A_{(n)k}^2
(Hax)^{-2\Lambda_k/H},
\label{series}
\end{eqnarray}
where
\begin{equation}
A_{(n)k}=\int_{-\infty}^\infty y^ne^{-v(y)}\Phi_k(y)dy
\label{ak}
\end{equation}
and we have used Eqs.~(\ref{Pi}), (\ref{ta}), (\ref{p2}) and the
orthonormality of the functions $\Phi_k(y)$. The mode functions
$\Phi_k(y)$ are even functions of $y$ for $k$ even and odd functions
of $y$ for $k$ odd ($\Phi_k(y)$ has $k$ nodes), and it is clear from
Eq.~(\ref{ak}) that
$A_{(n)k}$ are non-zero only when $n$ and $k$ are both even or both odd.
The first term in the series (\ref{series}) is
\begin{equation}
N^{-1}A_{(n)0}^2=\langle y^n\rangle^2,
\end{equation}
and thus the reduced two-point function,
\begin{equation}
c_n(x)=\langle y_1^ny_2^n\rangle-\langle y^n\rangle^2,
\end{equation}
is given by the same series starting with $k=1$. The asymptotic
behavior of $c_n(x)$ at large $x$ is determined by the first term in
that series. For odd values of $n$,
\begin{equation}
c_n(x)\approx N^{-1}A_{(n)1}^2(Hax)^{-2\Lambda_1/H}.
\end{equation}
For even values of $n$, $A_{(n)1}=0$ and
\begin{equation}
c_n(x)\approx N^{-1}A_{(n)2}^2(Hax)^{-2\Lambda_2/H}.
\label{xi2}
\end{equation}
Eqs.~(\ref{11}), (\ref{12}) of Section III follow immediately from these
relations.
Quite similarly, the three-point function can be expressed as
\begin{equation}
\langle y_1^4y_2^4y_3^4\rangle =\int dy_1dy_2dy_3 y_1^4y_2^4y_3^4
P_3[y_1({\bf x_1}),y_2({\bf x_2}),y_3({\bf x_3})].
\label{tpf}
\end{equation}
Assuming that ${\bf x_1},{\bf x_2},{\bf x_3}$ form an equilateral
triangle of side $ax$, the probability distribution $P_3$ can be
written as
\begin{equation}
P_3[y_1,y_2,y_3]=\int\Pi(y_1,0|y_a,t_a)
\Pi(y_2,0|y_a,t_a)\Pi(y_3,0|y_a,t_a) P_0(y_a)dy_a,
\label{p3}
\end{equation}
Combined with equations (\ref{tpf}),(\ref{Pi}),(\ref{ta}), this gives
\begin{equation}
\langle y_1^4y_2^4y_3^4\rangle=N^{-1}\sum_{k,l,m=0}^\infty A_{(4)k}A_{(4)l}
A_{(4)m} (Hax)^{-(\Lambda_1 +\Lambda_2 +\Lambda_3)/H}
\int\Phi_k(y)\Phi_l(y)\Phi_m(y) e^{v(y)}dy.
\end{equation}
Once again, it is easily understood that the reduced three-point
function (\ref{17}) is given by the same series with the summation
starting at $k,l,m=2$. The leading term at large $x$ is
\begin{equation}
\xi_3(x)\approx N^{-1}A_{(4)2}^3 C(Hax)^{-3\Lambda_2/H},
\label{xi3}
\end{equation}
where
\begin{equation}
C=\int_{-\infty}^\infty \Phi_2^3(y)e^{v(y)}dy.
\label{C}
\end{equation}
>From (\ref{xi2}) and (\ref{xi3}) we find that at large $x$
\begin{equation}
\xi_3(x)/\xi_2^{3/2}(x)\approx N^{1/2}C,
\end{equation}
independent of $x$.
|
\section{Introduction}
The aim of this talk is to describe a mathematics paper\cite{EJJ} written
by the same authors to a community of scientists
not necessarily specialized in the theory of fractals. We hope that
the concepts developed here can help to understand some issues which
have occurred in connection with the debate about the dimension of the
galaxy distribution.
We want to explain the
intuitive motivations of our definitions and results in the context of
actual physical problems. These motivations will be underlined by
simple but relevant examples.
The observation and study of fractal sets, {\it i.e.}, sets with
possibly non-integer dimension, is of course widespread. The present
work deals with two new aspects of fractal sets:
\begin{list}{}{}
\item{$\bullet$\ }Can the presence of voids be used to say something about
the dimension of a set?
\item{$\bullet$\ }How do voids which are almost empty account for the
dimension of a set?
\end{list}
We will give some answers in the case of sets which
are not too
wild, namely satisfying a certain doubling condition.
Our interest in the question of voids in fractals (porosity) was
raised by questions about the fractal dimension of galaxy
distributions. In view of the heated debate in this subject, see {\it e.g.}
\cite{D,PMS},
it seems adequate to provide as many analytical tools as possible with
which the fractal dimension can be estimated. This paper provides a new
such tool, namely the {\em porosity of the measure on the set}. The
general idea is that sets with large voids must have small dimension.
These voids, {\it e.g.}~in galaxy distributions, are taken as indicators of
small dimension in a sense we make precise below, and our theory
allows a systematic way to disregard occasional points (galaxies)
inside a (large) void. We present here a formalism which is tailored
for this situation, by presenting algorithms for measures rather than
for their supports.
\section{Measures and sets}
The main idea\cite{M1,S} describing the relation between the porosity and
the dimension goes about as follows, and we present some intuitive
examples which can guide the reader unfamiliar with this problem. Take
the well-known middle third Cantor set $C$. Clearly, if we consider
the voids in this set, every point in $C$ is close to a relatively
large void (namely to the middle third which has been taken out).
We will
give a more precise definition below.
Another way to view porosity, which is closer to the actual definition
is as follows: Suppose we have found a void of diameter $\rho$ inside
some minimal ball of radius $r'$ around a given point of the
fractal. Then the question is: How big a radius we have to take in order to
see for the first time a bigger void than the one we have
already seen? See also \fig(lacundivp.ps) below.
It is clear that if we take out
more of the middle, we make the dimension of the set
smaller, since the dimension depends in a well-known fashion on the length
ratio of voids to non-voids, namely, if we remove an interval of length
$\frac{k-2}k$ from the middle, {\it i.e.}, leave two intervals of length $\frac 1k$,
then the dimension is (in $\real$)
$\log 2/\log k$. For example for the middle third Cantor
set the dimension is $\log 2/\log 3$.
Thus, {\em large voids imply
small dimension}. However, the contrary is {\em not} true, as we shall
explain in Example 3 below.
One can construct a sequence of regular fractals, all of the
same dimension, but with porosity decreasing to 0.
Thus, {\em a set can have small dimension
without any porosity}. It is this aspect which is connected with the
controversy about the dimension of the galaxy distribution.
The requirement of obtaining information about
experimentally measurable objects leads us to consider measures, or
mass distributions, rather than sets.
This issue was addressed earlier\cite{ER} in the context of dimension
measurements. For example, one can compute the dimension of the
{\em support} of a set, {\it i.e.}, of the complement of the open sets of zero
measure. But, as explained in\cite{ER}, the so-called correlation
dimension which is based on the {\em mass} in balls seems to be the natural
quantity for questions of experimental nature and is commonly used in
the Grassberger-Procaccia method\cite{GP}.
Therefore, we shall study here the porosity of a {\em measure}, and
not only the porosity of the support of the measure
\cite{ER}.
We then show that large porosity implies a non-trivial upper bound on
the dimension (in fact on all multi-fractal dimensions
$D_q$, $q>1$). Finally, we explain how porosity is estimated for a given set
of experimental points.
\section{Porosities of measures}
Let $\mu$ be a probability measure on
$\real^n$. We define for $x\in \real^n$ and $r,\epsilon >0$:
\begin{eqnarray*}
{\rm por}&(\mu,x,r,\varepsilon)\,
=\,\sup\{{p}\ge 0~: {\text{there is }z\in\real^n
\text{ such that }} \\
&B(z,p r)\subset B(x,r)\text{ and }\mu(B(z,p
r))\le\varepsilon\mu(B(x,r))\}~.
\end{eqnarray*}
In other words, we consider the ball $B(x,r)$ of radius $r$ centered at
$x\in\real^n$ and observe the mass $\mu(B(x,r))$ contained in it.
We now look for
the largest ball of radius $p\cdot r$
(fully contained in $B(x,r)$) around a point
$z$ such that the mass of that ball does not exceed $\epsilon
$ times the mass $\mu(B(x,r))$ of $B(x,r)$.\footnote[1]{The conventional porosity asks for the largest ball
in $B(x,r)$ which does not contain {\em any} point of the set in question;
see also below.}
See \fig(fig1p.ps).
One then defines
\begin{equation}\label{eq@pmx}
{\rm por}(\mu,x)\,=\,\lim_{\epsilon \downarrow0}\,\liminf_{r\downarrow0}
{\rm por}(\mu,x,r,\epsilon )~,
\end{equation}
and finally
\begin{equation}\label{eq@intropor}
{\rm por}(\mu)\,=\,\inf\{s~:~ {\rm por}(\mu,x)\le s \text{ for }\mu\text{-almost
all }x\in\real^n\}~.
\end{equation}
Note that when $x$ is not in the support of the measure, the
definition is not very interesting since for small enough $r$ the
measure of $\mu(B(x,r))$ is zero and then any ball in it is also empty.
\begin{@MyDefinition}The quantity ${\rm por}(\mu)$ is called the porosity
of the measure $\mu$.
\end{@MyDefinition}
\FIGUREWITHTEX{fig1p.ps}{fig1.tex}{100}{1}{0}{-0.9}{1}{The ball
$B(z,p r)$ inside $B(x,r)$.}
It is not difficult to see that the definition of the {\em porosity of a
set} (see Definition \ref{definition@defpor} below) amounts to using
Eq.\equ(pmx) with the
limits taken in the
opposite order, that is,
\be{pmx2}
{\rm por}(\spt(\mu),x)\,=\,\liminf_{r\downarrow0}\,\lim_{\epsilon \downarrow0}
{\rm por}(\mu,x,r,\epsilon )~.
\ee
Because the central point in $B(x,r)$ is in the support $\spt(\mu)$
and hence occupied when we compute ${\rm por}(\spt(\mu),x)$, one finds that
${\rm por}(\spt(\mu))\le\frac12$. One can also show, using density
arguments, that
${\rm por}(\mu)\le \frac12$. We also note that ${\rm por}$
is determined {\em first} with ``dust'' of relative weight $\epsilon $ and only
then $\epsilon$ is taken to 0.
The two porosities we consider satisfy clearly
${\rm por}(\spt(\mu))\le{\rm por}(\mu)$. In other words the porosity of a
measure is larger than that of the support of the measure, precisely
because the former neglects occasional dust.
\LIKEREMARK{Example 1}
Let $\delta_0$ be the Dirac measure at the origin, that is,
$\delta_0(A)=1$ if $0\in A$ and $\delta_0(A)=0$ if $0\notin A$.
Let $\mu$ be the sum of $\delta_0$ and the Lebesgue measure ${\cal L}^n$
restricted to $B(0,1)$, that is, $\mu=C(\delta_0+{\cal L}^n|_{B(0,1)})$
where $C$ is the normalization constant. Clearly
${\rm por}(\mu,0)=\frac 12$ and ${\rm por}(\mu,x)=0$ for all $x\ne0$ with $|x|<1$.
Thus ${\rm por}(\mu)=\frac 12$. However, ${\rm por}(\spt(\mu))={\rm por}(B(0,1))=0$.
\LIKEREMARK{Example 2}We next want to argue that voids are accounted
for in a more
reasonable way in the measure theoretic definition of porosity.
To illustrate this with a concrete example,
consider the celebrated middle third Cantor set which is
obtained by starting with the interval $[0,1]$ and taking out the open
interval $(\frac13,\frac23)$. Then each of the remaining two intervals
is divided into
three pieces and the middle one (of length $\frac19$) is discarded. Going on
recursively (and indefinitely) in this fashion, we get the middle
third Cantor set. Its
dimension is $d\equiv \frac{\log2}{\log3}$. Clearly, this set $C$ has voids
and its porosity equals in fact $\frac14$.
We next set $C_x=\{ y~:~ y=x+z, z\in C\}$, in other words, $C_x$ is
the translate
of $C$ by $x$. Clearly each $C_x$ has again dimension $d$. From
the general theory of fractals\cite{F,M2} we get that any countable union of
such sets has still dimension $d$. In particular, we can enumerate the
rationals in $[0,1]$, for example calling them $x_j$, $j=1,2,\dots$ and
construct then the set
$$
D\,=\,\bigcup_{j=1}^\infty C_{x_i}~.
$$
From what we said before, and from the way we constructed it, the set
$D$ has dimension $d<1$ and {\em no voids}. Thus, we see that a {\em
set} can have small dimension and no voids.
The example we have just given does {\em not work} in the case of porosity
of measures and this is one of the reasons why the porosity of
measures is a more useful concept than that of sets.
However, we will construct regular fractals in Example 3 where the
porosity of the measure is arbitrarily small but the dimension is
always $\frac12$.
In fact, one shows easily, see\cite{EJJ},
that the following measure has porosity $\frac 12$.
We construct a measure on the set $D$ by giving successively lower
weight to the translates of $C$. Let $\mu $ be the usual measure
associated with the Cantor set $C$, {\it i.e.}, the measure which gives equal
weight to all the pieces in the recursive construction. Then we define
$$
\nu\,=\,\sum _{i=1}^\infty 2^{-i} \mu_{x_i}~,
$$
where $\mu_{x}$ is the translate of the measure $\mu$ by $x$. Clearly,
the support of this measure is at least all of the interval $[0,1]$
(with some overhangs from the translation)
and has therefore no voids. But the porosity ${\rm por}(\nu)$, as defined in
Eq.\equ(intropor), is strictly positive. The reason for this is that if
we consider a void of the original set $C$ and look for a point in one
of the $C_{x_i}$ very
close to the boundary of this void, then we must take in general a
high index $i$ to find such a point. But then the associated
weight is smaller than $2^{-i}$,
and if $i$ is large enough, then this is smaller than any $\epsilon $
which was given in the definition of Eq.\equ(pmx). Therefore, the set
$C_{x_i}$ is not counted in this consideration, and the measure
will have the same porosity as $C$ itself.
Having described the definitions, we can now ask more precisely our question
about the relation between the porosity of the measure $\mu$ and the
packing dimension of the same measure. Our general aim is to show the
following
\begin{@MyConjecture}If the porosity of $\mu$ is large, then the
packing dimension of $\mu$ is smaller than the dimension $n$ of the
ambient space.
\end{@MyConjecture}
Our results will fall somewhat short of this conjecture.
In order to be able to formulate a positive result, we need the
following concept:
\begin{@MyDefinition}The probability measure $\mu$ on $\real^n$
satisfies the local doubling condition at $x$ if
\begin{equation}\label{eq@20}
\limsup_{r\downarrow0} {\m(x,2r)\over \m(x,r)}\,<\,\infty ~.
\end{equation}
It satisfies the (global) doubling condition if Eq.\equ(20) holds for
$\mu$-almost
every $x$.
\end{@MyDefinition}
The bound need not be uniform in those $x$.
Note that the doubling condition is
also implied by the stronger condition
\be{stronger}
0\,<\, a r^s \,\le\, \mu(B(x,r))\,\le\,b r^s \,<\,\infty ~.
\ee
In physics, it is generally assumed that the stronger condition
\equ(stronger) holds.
See below for the relevance of these conditions in Nature.
Our main result is the following
\begin{@MyTheorem}\label{theorem@main}There is a function $\Delta_n$ defined for
$p\in(0,1/2)$ with values in $[0,1]$ and satisfying
$$
\lim_{p\to{1\over 2}} \Delta_n(p)\,=\,1~,
$$
such that
if a Borel probability measure $\mu$ on
$\real^n$ satisfies the global doubling condition
then
\begin{equation}\label{eq@ineq}
\dimH(\mu)\,\le\,\dimp(\mu)\,\le\, n-\Delta_n({\rm por}(\mu))~.
\end{equation}
\end{@MyTheorem}
Here, $\dimH(\mu)$ is the Hausdorff dimension of the measure and
$\dimp(\mu)$ is its packing dimension. The inequality among those two is
obvious from their definition:
\begin{eqnarray*}
\underline
d(\mu,x)&\,=\,&\liminf_{r\downarrow0}{\log\mu(B(x,r))\over\log r}~,\\
\overline d(\mu,x)&\,=&\,\limsup_{r\downarrow0}{\log
\mu(B(x,r))\over\log r}~,\\
\dimH(\mu)&\,=\,&\sup\{s\ge0~:~
\underline d(\mu,x)\ge s\text{ for }\mu
\text{-almost all }x\in \real^n\}~,\\
\dimp(\mu)&\,=\,&\sup\{s\ge0~:~
\overline d(\mu,x)\ge s\text{ for }\mu
\text{-almost all }x\in\real^n\}~.
\end{eqnarray*}
There is an explicit lower bound for the function
$\Delta_n$ in\cite{S}:
$$
\Delta_n(p)\,\ge\,\max\{1 - {c_n\over \log(\frac1{1-2p})},0\}~,
$$
where $c_n>0$ is a constant depending only on $n$.
According to Theorem \ref{theorem@main} if the porosity of a measure $\mu$ which satisfies
the doubling condition is close to $\frac12$, then the packing
dimension of $\mu$ is not much bigger than $n-1$.
\LIKEREMARK{Remark} For sets, a relation between porosity and dimension
has been established by Mattila\cite{M1} and Salli\cite{S} using the
following definition of porosity:
\begin{@MyDefinition}\label{definition@defpor}The porosity of a set $A\subset\real ^n$ at
a point $x\in\real ^n$ is defined by
$$
{\rm por}(A,x)\,=\,\liminf_{r\downarrow0}{{\rm por}(A,x,r)}~,
$$
where
$$
{\rm por}(A,x,r)=\sup\{ p\ge0:\,\text{ there is }z\in\real ^n \text{
such that } B(z,p r)\subset B(x,r)\setminus A\}.
$$
Here, $B(z,\alpha)$ is the closed ball with radius $\alpha$ and with center at
$z$.
The porosity of $A\subset\real ^n$ is
$$
{\rm por}(A)=\inf\{{\rm por}(A,x): x\in A\}~.
$$
\end{@MyDefinition}
We do not
know any example where
\begin{list}{}{}
\item{$\bullet$\ }The porosity of the measure is $\frac12$,
\item{$\bullet$\ }and the dimension of the measure is $n$ (in the ambient space
$\real^n$).
\end{list}
\noindent Of course, this would have to be a measure which violates
the doubling
condition.
The doubling condition is a bound on the amplitudes of the
fluctuations of the integrated density of the measure. The role of
this condition in experiments in galaxy distributions is somewhat
obscure.
But, for example, in\cite{MPST} the authors measured 2 periods of
density fluctuations (of about the same amplitude) and so in a very
weak sense, the doubling condition seems experimentally
satisfied.
Another example is given in \fig(doublingp.ps) where the
results of measuring the doubling condition for certain catalogs of
galaxies are shown.
\FIGUREWITHTEX{doublingp.ps}{null.tex}{1}{0.65}{0}{-0.65}{0}{Verification
of the doubling condition for several central points of two galaxy
catalogs. We thank M. Montuori and F. Sylos-Labini for these calculations.
The fluctuations seem quite bounded, except at short distances, where
there are problems because one runs out of data.}
For regular recursively constructed fractals the doubling condition
is always satisfied. Finally, we believe that fractals formed in
Nature by a physical law have not only the same dimension everywhere,
but also satisfy the doubling condition. The reason for this is that for
example an attractor looks everywhere similar because it is
created by a (smooth) physical law, which transports the structure of
the fractal around in space (making at most smooth coordinate changes
locally). This point of view has been advocated in\cite{ER}, and has
been rigorously verified for a few non-trivial examples of dynamical systems.
Another class of measurements takes the density
fluctuations of the mass itself as an indicator of the dimension of
the measure. This seems a mathematically inaccessible (and probably
wrong) criterion for dimension measurements. It might be that this
idea is a consequence of the regular oscillations one gets for regular
Cantor sets. The only case where a rigorous result is known is that of
integer dimension:
\begin{@MyTheorem}\label{theorem@Mar}(Marstrand) Let $s$ be a positive integer.
Suppose that there exists a Radon measure $\mu$ on $\real^n$ such that the
density
$$
\lim_{r\downarrow 0}{\mu(B(x,r))\over r^s}
$$
exists and is positive and finite in a set of positive $\mu$-measure. Then $s$
is an integer.
\end{@MyTheorem}
(For the proof see\cite{M2} Theorem 14.10.)
It is well-known that one cannot expect an inequality in the sense opposite
to the one stated in the theorem. That is, big voids imply small
dimension, but small
voids do not imply big dimension. This is illustrated by the
following example in $\real$, {\it i.e.}, in one dimension.
\LIKEREMARK{Example 3}We will construct a sequence of measures
$\mu^{(n)}$, all of dimension $\dimp(\mu^{(n)})={1\over 2}$ in $\real$ with
porosity
${\rm por}(\mu^{(n)})\le {1\over n}$. The set $A^{(n)}$ is a Cantor set obtained
recursively as
follows: Divide the interval $[0,1]$ into $n^2$ equal subintervals
and select $n$ of these subintervals, namely the $1^{\text {st}}$,
$n+1^{\text{st}}$, and so on.
The measure at this level of the construction
is obtained by giving the same weight ${1\over n}$ to each
subinterval of $A^{(n)}$. In other words, the unit measure is
uniformly distributed on the $n$ intervals constructed so far.
\FIGUREWITHTEX{lacundivp.ps}{lacundiv.tex}{1}{0.54}{0}{-0.65}{1}{The
oscillations of ${\rm por}(r)$ as a function of $r$ (on a logarithmic scale)
for the sets $A^{(n)}$, with $n=3,\dots,6$.}
Now repeat inductively the procedure for
each of the $n$ intervals, dividing it into $n^2$ equal pieces and
selecting each $n^{\text{th}}$ among them. Give each of these
intervals weight $\frac1{n^2}$.
\FIGUREWITHTEX{lacunnodivp.ps}{lacunnodiv.tex}{1}{0.54}{0}{-0.65}{1}{The
largest void $\rho$ as a function of $r$, measured from the left most
point in $A^{(n)}$
for $n=3,\dots,6$.}
Continuing indefinitely in this fashion, one obtains the Cantor set
$A^{(n)}$ and
the measure $\mu^{(n)}$ on it.
The dimension of this measure is $\dimp(\mu^{(n)})= \log( n) / \log
(n^2)=\frac12$, and it is not difficult to check that the porosity is
less than
$\frac1n$. (Since the gaps become smaller for larger $n$, see\cite{EJJ}.)
In this case, it is easy to compute numerically the porosity of the
sets $A^{(n)}$ when $n$ is not too large. In \fig(lacundivp.ps) we
show the quotient ${\rm por}(\mu,x,r,\epsilon =0)$ as a function of $r$
when $x$
is the leftmost point of $A^{(n)}$. Similar, but more irregular
pictures are obtained when one chooses another point $x$.
One can understand the origin of the oscillations by looking at
\fig(lacunnodivp.ps), where we plot the radius $\rho$ of the
largest empty interval as a function of $r$.
We see that $\rho $ grows linearly, until a point of the Cantor set is
hit, and then it stays constant until a bigger void is found.
Then \fig(lacundivp.ps) is obtained by dividing the values obtained
in \fig(lacunnodivp.ps) by $r$.
The proof of our main result is based on comparing the porosity of a
measure with the porosity of subsets with positive measure. For this,
we use the quantity $\beta (\mu)$ introduced in\cite{MM}:
$$
\beta(\mu)\,=\,\sup\{{\rm por}(A): A\text { is a Borel set with }\mu(A)>0\}~.
$$
The inequality $\beta (\mu)\le{\rm por}(\mu)$ holds for any Borel
probability measure $\mu$, but the converse inequality does not need to be
true. We show that it holds when the measure $\mu$ satisfies the
doubling condition. We have shown in\cite{EJJ} that the doubling
condition implies
$\beta
(\mu)={\rm por}(\mu)$. Together with the results of\cite{S} it implies
our main bound on the dimension (see Theorem \ref{theorem@main}).
We also showed that there
are measures violating the doubling condition for which $\beta
(\mu)\ne{\rm por}(\mu)$. In the case of measures on the line $\real $, {\it
i.e.,} in 1 dimension, we
also show that a somewhat weaker condition than the doubling condition
implies Theorem \ref{theorem@main}.
|
\section{High Redshift Structure}
The most puzzling aspect of the distant dropout galaxies is the
appearance of sharp peaks in the redshift distribution
(Adelberger{\sl~et\,al.~} 1998, Steidel 1998), resembling the situation at low
redshift, implying at face value little growth of structure. A
conventional interpretation of these peaks resorts to ``bias''
(Wechsler {\sl~et\,al.~} 1998) which has come to represent a flexible
translation between the observed structure and the relatively smooth
mass distribution of standard simulations, so that the observed peaks
are interpreted as rare events destined to become massive clusters by
today (Wechsler{\sl~et\,al.~} 1998). High biases have been claimed for the
Lyman-break population on the basis of the amplitude of small scale
clustering at $z\sim3$ (Adelberger{\sl~et\,al.~} 1997,
Giavalisco{\sl~et\,al.~} 1998). The occurrence of such peaks is enhanced with a
steep spectrum by a reduction of high frequency `noise' (Wechsler et
al 1998) but this gain is offset if the steepness is attributed to low
$\Omega$, since then a given redshift bin corresponds to a larger
volume, and hence a greater proper density contrast (Wechsler et al
1998). We may regard the existence of regular spikes at low and high
redshift as evidence for a revision in our understanding of large
scale structure, indicating perhaps that initial density fluctuations
are not Gaussian distributed such as may be implied by the observed
lack of evolution of the number density of X-ray selected clusters
(Rosati {\sl~et\,al.~} 1998). The baryon isocurvature model (Peebles 1997) more
naturally accommodates both the early formation and the frequent
non-Gaussian occurrence of high density regions (Peebles 1998a,b).
We analyze the fields of Lyman-Break galaxies of Steidel (1998) and
Adelberger(1998). These include 4 fields of $\sim 100-200$ galaxies
and a smaller sample of redshifts in the Hubble Deep Field
direction. The fields are $\sim10$Mpc in width by $\sim 400$Mpc in the
redshift direction and include over 600 redshifts in the range
$2.5<z<3.5$ histogrammed in bins of $\Delta z=0.04$. In Figure~1 we
plot the pair counts and correlation function, assuming that galaxies
are evenly distributed within the narrow redshift bins. A clear excess
is apparent on a large scale, corresponding to a preference for
separations of $\Delta z=0.22$. A pair excess is also seen at twice
this separation (Fig~1) indicating phase coherence along
the redshift direction. This behaviour is an obvious consequence of
the regular peaked structure visible in the redshift histograms of
four of the five fields (Fig~1). In a bin of 10Mpc, the number of pairs at the
peak is $\sim 1220$ compared with an expected $\sim 1060$,
representing a $4.6\sigma$ departure from random (Fig~1).
The observed redshift interval may be related to comoving scale
by the usual formula for a universe with negligible radiation
energy-density (Peebles 1993),
\begin{eqnarray}\label{eq:deltaw}
\Delta w &=& 3000\; h^{-1} \;{\rm Mpc}\; \int_z^{z+\Delta z} {dz\over E(z)}
\nonumber\\
&\simeq& 3000\; h^{-1} \;{\rm Mpc}\; {\Delta z\over E(z)}
\end{eqnarray}
with
\begin{equation}
\label{eq:Ez}
E(z) = \sqrt{
\Omega (1+z)^3 + \Omega_\Lambda + (1-\Omega-\Omega_\Lambda)(1+z)^2}.
\end{equation}
Whereas standard candles at moderate redshift (SNIa) measure
$\sim(\Omega - \Omega_\Lambda)$, and CMB anisotropies measure
$\sim(\Omega + \Omega_\Lambda)$, these observations considered here
measure $\sim E(3)$ or $\sim (3\Omega-\Omega_\Lambda)$ which lies
between these two locii, thus adding complimentary information.
With this, the $\Delta z=0.22$ scale of the peak corresponds to
$85\;h^{-1}{\rm Mpc}$ for $\Omega=1$, doubling to $170\;h^{-1}{\rm
Mpc}$ for $\Omega=0$. Flat cosmologies with a positive $\Lambda$ fall
between these limits. If we constrain $\Delta w = 130\; h^{-1}{\rm
Mpc}$ we find $48\Omega-15\Omega_\Lambda \simeq 10.5$. This gives
$\Omega=0.2$ for an open universe ($\Omega_\Lambda\equiv0$) or
$\Omega=0.4$ for a flat universe ($\Omega+\Omega_\Lambda\equiv1$),
with an uncertainty of only $0.1$ in $\Omega_m$, given by the 15\%
width of the excess correlation (Figure~1). This scale, if borne out
by subsequent redshift data, is certainly consistent with the flat
model preferred by SNIa data with $\Omega_m^{flat}=0.3$ (Perlmutter {\sl~et\,al.~}
1998,1999).
\section {Excess Power and the CMB}
One can imagine two broad classes of physical mechanisms that might
be responsible for the excess power required to fit this data. The power might be a truly primordial feature in the power
spectrum. In inflationary scenarios, for example, this could be
generated by the proliferation of super-horizon bubbles in a suitably
conspiratorial inflaton potential (La 1991, Occhinero \& Amendola
1994). On the other hand, excess power could be due to causal
microphysics in the universe after the 130~$\;h^{-1}{\rm Mpc}$ scales
enter the horizon. In other words, the power could be added by the
transfer function. A high baryon density naturally imparts large scale
``bumps and wiggles'' to the power spectrum (Peebles 1998a,b,
Eisenstein{\sl~et\,al.~} 1997, Meiksin{\sl~et\,al.~} 1998) in particular if matter
density fluctuations are created at the expense of radiation (Peebles
1998a,b). An even more radical possibility is that the universe is
topologically compact, and that we are seeing evenly-spaced copies of
a small universe with an extent of only $\sim130\;h^{-1}{\rm Mpc}$
although this scale seems too small to accommodate the unique and
relatively distant (z=0.18) cluster A1689 (Gott 1980).
In light of this it is interesting to explore the implications of any
excess power on CMB temperature fluctuations. Eisenstein{\sl~et\,al.~} (1997)
has examined in what way conventional adiabatic models maybe stretched
to match the power spectrum of Broadhurst{\sl~et\,al.~} (1990), demonstrating
that such models do not naturally account for a scale of
130$\;h^{-1}{\rm Mpc}$ in the mass distribution.
Our approach is to simply add power in the primordial spectrum at a
fixed wave number, using a modified version of CMBFAST code (Seljak \&
Zaldarriaga 1999), to simulate temperature fluctuation spectra. A
primordial spike of power will effect the CMB directly through the
projection of three-dimensional features. A narrow band of power is
added, $\Delta^2(k) = k^3 P(k)/(2\pi^2)$. The amplitude is set so that
$\int d\ln{k}\; \Delta^2(k) \simeq 0.1$, the value of the correlation
function at the peak and harmonics equivalent to at 30\% density
contrast. We ignore bias, which may conceivably be large,
lowering the peak amplitude; we also ignore the possibility of
non-Gaussian fluctuations which would certainly diminish the
angle-averaged power.
If the value of the power spectrum at the $k=0.05\; h^{-1}{\rm Mpc}$
peak is already large, as with COBE-normalized $\Lambda$ models, an
additional peak of the above strength has negligible effect. However,
COBE-normalized models with $\Omega=1$ and $\Lambda=0$ have
comparatively low power at this scale, hence, the effect on the final
$C_\ell$ and $P(k)$ is significant. Of course, the precise width and
location of the peak, not yet at all well-constrained by the CMB data,
also affects the relative power in the peak versus the underlying
``smooth power-law'' spectrum. This is illustrated in Figure~2. The
spike is seen to raise the amplitude of the first Doppler peak by
nearly a factor of two for the sCDM model.
Note, although it is premature to interpret the claims of the various
CMB experiments for and excess of power at $C_\ell=200$ (Netterfield
et al 1997, Tegmark 1998) without proper treatment of the covariance
matrix of errors and foreground subtraction, the indications of a
large amplitude for this peak are not inconsistent with the degree of
boosting (Fig.~2). Others, Gawiser and Silk 1998, have noted that the
whole aggregate of current $C_\ell$ and $P(k)$ data is extremely
well-fit---much better than the fit to standard models---by an
adiabatic inflationary power spectrum with sharp bump like that
considered here.
\section {Conclusions}
Two interesting results emerge from the above comparison of structure
in the local and distant pencil beam data. Firstly both samples of
galaxies show regular large scale power confined to a narrow range of
frequency. Secondly, by matching these scales we obtain values of
$\Omega$ and $\Lambda$ in good agreement with the SNIa claims. These
findings may of course be regarded as remarkable coincidences, unlikely
to occur by chance in a clumpy galaxy distribution (Kaiser \& Peacock
1991).
Distinguishing between the open and flat solutions found here requires
pencil beam data at a third redshift. Optimally this redshift turns
out to be convenient for observation, $z\sim0.7$, where the ratio of
$\Delta z$ between the open $\Omega=0.2$ and $\Omega_m^{flat}=0.4$
cases is maximal, differing by 13\%, and may be explored with
existing data.
Since locally the excess power at 130$\;h^{-1}{\rm Mpc}$ is most
prominent in cluster selected samples, we may conclude that the peaks
in the high-$z$ sample correspond to proto-clusters at $z\sim 3$. This
conclusion is independent of the high-bias interpretation which also
regards the spikes as young rich clusters.
Understanding the physics behind these spikes is complicated by a
conspiracy of scales. We have seen in Fig.~2 that the
$k\simeq0.05\;{\rm Mpc}/h$ spike produces a feature
in the CMB power spectrum at nearly exactly the position of the first
doppler peak. In the simple model presented here---a primordial spike at
this position---this is merely a coincidence, however unlikely.
That these scales are so closely matched seems yet more unlikely in the
light of another fact, that the {\em required} peak in the matter power
spectrum is at nearly the position of the {\em expected} peak due to the
passage from radiation to matter-domination.
If the acoustic peaks in the CMB power spectrum are instead a
signature of physics at a somewhat later epoch: the sound horizon at
recombination (set by $c/\sqrt{3}$) then that these scales are
nearly coincident (within an order of magnitude or so of 100
$\;h^{-1}{\rm Mpc}$) is a consequence of the particular values of the
cosmological and physical parameters. Thus, if we change the initial
conditions to increase the amplitude of the matter power spectrum near
the equality scale, we also increase the CMB temperature power
spectrum at roughly the scale of the first acoustic peak. That is, in
neither case do we add an unexpected peak, but merely increase the
amplitude and ``sharpness'' of an expected one.
Of course, the inexplicable coincidence is that the feature in the
power spectrum appears close to the expected matter-radiation equality
peak, but with an amplitude much too large to be explained by standard
theories. In addition, we must also note that merely adding a peak to
the mean power spectrum cannot account for the nearly periodic
structure observed in 1D. A real theoretical explanation must account
for the complicated non-Gaussian properties of this distribution.
The excess 1D power identified by Broadhurst{\sl~et\,al.~} (1990) will soon be
subject to easy dismissal with large 3D redshift surveys. However,
whether or not it transpires that universal large scale coherence
exists, the test proposed here is still viable in principle using the
general predicted turnover of the 3D power spectrum, requiring larger
volume redshift surveys. In particular comparisons of local and
distant clusters over a range of redshift will be particularly useful
given their sharply peaked and high amplitude power spectrum..
\acknowledgments
We thank Richard Ellis, Alex Szalay, Gigi Guzzo, Alvio Renzini, Eric
Gawiser, Saul Perlmutter and Martin White for useful conversations.
\fontsize{10}{14pt}\selectfont
|
\section{Introduction}
Spread options have become increasingly important. They give the holders the
right to call or put the spread value of two underlying assets against a
predetermined parameter $K$ as the strike price. In particular, the spread
options reduce to the so-called exchange options when the predetermined
strike price $K$ is set to zero. Spread options and exchange options can be
viewed as options to exchange one underlying asset for another with respect
to the strike price. They are used in many situations. One typical example
is that the option holder is interested in exchanging one commodity for
another commodity. For instance, in oil industries, the prices of crude oil
and refined oil differ from each other, and both prices are fluctuating
considerably in response to the weather, regional stabilities of world oil
production centers, and other human and natural parameters. Oil companies
may deal with the situations of price fluctuations using the spread options
or exchange options. Spread and exchange options have been of considerable
interests to both practitioners and theoretical researchers\cite{marg,garman,tan}.
For the exchange option, a closed form solution for its price is available.
The valuation of exchange option was first
studied by Margrabe\cite{marg}, based
on the option pricing theory of Black and Scholes\cite{black}, and Merton\cite{merton1,merton2}.
The derivation of Margrabe is a PDE approach\cite{marg}.
However, in Margrabe's
derivation, the risk-free rate $r$ is assumed to be a constant, which is far
from reality. It is of great interest to both practitioners and theoretical
researchers to investigate whether closed form solution exists when the
interest rate is modeled with stochastic term structure.
This paper investigates how stochastic interest rate will affect the
exchange option pricing. The closed form solution for the exchange option's
price is given when we assume a very general stochastic process for the
interest rate, which includes Vasicek model\cite{vasicek},
CIR model\cite{cir}, affine term
structure models\cite{kan}
and other interest rate models as special cases. To our
knowledge, this is the first time to provide the closed form solution for
exchange option pricing while stochastic interest rate is taken into account.
We also argue that to price a European style spread option of general strike
price $K$, one may use our closed form solution as a control variate to
reduce variance of simulation when doing Monte Carlo pricing\cite{phelim}
for stochastic interest rate. The closed
form result presented here shall be of interest to both theoretician and
practitioners.
In our discussion below, we assume an exchange economy populated by
risk-averse agents with increasing preferences, and all economic activities
take place in the time interval $[0,T]$. All the possible outcomes of
this economy is denoted by a measurable space $(\Omega ,\mathcal{F})$ where $%
\Omega $ is the set of all possible states and $\mathcal{F}$ is a sigma
algebra of subsets of $\Omega .$ Information arrival in this economy is
described by a filtration $\{\mathcal{F}_{t};0\leq t\leq T\}$ with $\mathcal{%
F}_{T}=\mathcal{F},$ and the agents belief is modelled by a probability
measure $P$ defined on $(\Omega ,\mathcal{F}).$
In following sections, we will first rederive the closed form solution of
exchange option using the risk-neutral martingale approach. It is shown that
this provides identical result to the one given by Margrabe using partial
differential equation. Then, we discuss, within the framework of martingale
measure, how to price the exchange option when interest rate is stochastic.
It is shown that our result is valid for most general term structure, only
the correlation coefficients between the interest rate and the underlying
assets will affect the option price.
\section{Exchange Options}
Exchange options can be defined by using two underlying assets or
commodities which are closely related. This correlation between the two
assets or commodities results from demand substitution or the potential for
transformation. In general, an exchange option has the following payoff:
\begin{equation}
\max \{\lambda (S_{1}(T)-S_{2}(T)),0\}=\left[ \lambda \left(
S_{1}(T)-S_{2}(T)\right) \right] ^{+} \label{a1}
\end{equation}
where $\lambda =1$ for a call and $\lambda =-1$ for a put, and $S_{1}(T)$
and $S_{2}(T)$ are the underlying asset prices at maturity $T$.
\subsection{Constant interest rate}
Let us first review the case of constant interest rate. The pricing closed
form was given by Margrabe via PDE method. In the following, we give a short
review within the framework of risk neutral
martingales\cite{cox, duffie,har1,har2}.
Part of the results will be used in the section of discussing stochastic interest rate case.
Suppose that in the physical probability space $(\Omega ,\mathcal{F},P)$,
the prices of two underlying assets for an exchange option follow the
geometric Brownian motions, that is,
\begin{equation}
\frac{dS_{i}(t)}{S_{i}(t)}=\mu _{i}dt+\sigma _{i}dW_{i}(t),\qquad i=1,\text{
}2 \label{ee5}
\end{equation}
where $\mu _{1},$ $\mu _{1},$ $\sigma _{1},$ $\sigma _{2},$ and $\rho
_{12}\equiv Corr[dW_{1},dW_{2}]$ are all constants. The price $%
C(t,S_{1}(t),S_{2}(t))$ of a (European) call exchange option at time $t$ is
then given by
\begin{equation}
C(t,S_{1}(t),S_{2}(t))=E_{t}^{Q}\left[ e^{-\int_{t}^{T}r(s)ds}\ \left[
S_{1}(T)-S_{2}(T)\right] ^{+}\right] \label{ee6}
\end{equation}
where $Q$ is the corresponding equivalent martingale measure. We should note
here that these two processes in (\ref{ee5}) are defined in the physical
probability space $(\Omega ,\mathcal{F},P),$ however, the general valuation
formula in (\ref{ee6}) is derived in the risk-neutral probability space $%
(\Omega ,\mathcal{F},Q).$ The relationship between the risk-neutral
probability space and the physical probability space is the standard one,
which is described by the Girsanov transformation\cite{duffie}.
If the risk-free rate $r$ is constant, then the closed form formula for
pricing the exchange option was first derived by Margrabe (1978) using the
partial differential equation approach. However, we can also compute the
expectation value in the risk neutral space, and the exchange option price
will follow. It is shown that the approach gives the pricing formula
identical to the one derived by Margrabe. In this case, we assume that there
is no dividend paying during the option's life. Since the price processes of
the two underlying assets are governed by (\ref{ee5}), under the equivalent
martingale measure $Q$ with constant risk-free rate $r$, we then have
\begin{equation}
\left[ \left.
\begin{array}{c}
\log [S_{1}(T)] \\
\log [S_{2}(T)]
\end{array}
\right| \mathcal{F}_{t}\right] \stackrel{D}{\thicksim }\mathcal{N}\left(
\left[
\begin{array}{c}
A_{1} \\
A_{2}
\end{array}
\right] ,\left[
\begin{array}{cc}
\nu _{1}^{2} & \rho _{12}\nu _{1}\nu _{2} \\
\rho _{12}\nu _{1}\nu _{2} & \nu _{2}^{2}
\end{array}
\right] \right) \label{qq3}
\end{equation}
where
\begin{equation}
A_{i}=\log [S_{i}(t)]+(r-\frac{\sigma _{i}^{2}}{2})(T-t),\qquad \nu
_{i}^{2}=\sigma _{i}^{2}(T-t),\qquad i=1,\text{ }2. \label{qq4}
\end{equation}
Note that we can write
\[
S_{1}(T)-S_{2}(T)=e^{A_{1}+\sigma _{1}\sqrt{T-t}Z_{1}}\left\{
1-e^{A_{2}-A_{1}+\sigma _{2}\sqrt{T-t}Z_{2}-\sigma _{1}\sqrt{T-t}%
Z_{1}}\right\}
\]
where
\[
\left[
\begin{array}{c}
Z_{1} \\
Z_{2}
\end{array}
\right] \stackrel{D}{\thicksim }\mathcal{N}\left( \left[
\begin{array}{c}
0 \\
0
\end{array}
\right] ,\left[
\begin{array}{ll}
1 & \rho _{12} \\
\rho _{12} & 1
\end{array}
\right] \right) .
\]
Hence,
\[
S_{1}(T)\geq S_{2}(T)\Leftrightarrow Z_{3}\geq m
\]
where
\begin{eqnarray*}
Z_{3} &\equiv &\frac{\sigma _{1}Z_{1}-\sigma _{2}Z_{2}}{\sqrt{(\sigma
_{1}^{2}+\sigma _{2}^{2}-2\rho _{12}\sigma _{1}\sigma _{2})}}, \\
m &\equiv &\frac{A_{2}-A_{1}}{\sqrt{(\sigma _{1}^{2}+\sigma _{2}^{2}-2\rho
_{12}\sigma _{1}\sigma _{2})(T-t)}},
\end{eqnarray*}
and
\[
\left[
\begin{array}{c}
Z_{1} \\
Z_{3}
\end{array}
\right] \stackrel{D}{\thicksim }\mathcal{N}\left( \left[
\begin{array}{c}
0 \\
0
\end{array}
\right] ,\left[
\begin{array}{ll}
1 & \eta \\
\eta & 1
\end{array}
\right] \right)
\]
with
\[
\eta \equiv \frac{\sigma _{1}-\sigma _{2}\rho _{12}}{\sqrt{\sigma
_{1}^{2}+\sigma _{2}^{2}-2\rho _{12}\sigma _{1}\sigma _{2}}}.
\]
So,
\begin{eqnarray}
&&E_{t}^{Q}\left[ \left[ S_{1}(T)-S_{2}(T)\right] ^{+}\right] \nonumber \\
&=&\int_{-\infty }^{+\infty }\int_{m}^{+\infty }e^{A_{1}+\sigma _{1}\sqrt{T-t%
}x}p(x,y,\eta )dydx-\int_{-\infty }^{+\infty }\int_{m}^{+\infty
}e^{A_{2}+\sigma _{1}\sqrt{T-t}x+by}p(x,y,\eta )dydx \nonumber \\
&\equiv &I_{1}-I_{2} \label{ii}
\end{eqnarray}
where
\begin{eqnarray*}
b &\equiv &\sqrt{(\sigma _{1}^{2}+\sigma _{2}^{2}-2\rho _{12}\sigma
_{1}\sigma _{2})(T-t)}, \\
p(x,y,\eta ) &\equiv &\frac{1}{2\pi \sqrt{1-\eta ^{2}}}e^{-\frac{1}{2(1-\eta
^{2})}(x^{2}-2\eta xy+y^{2})}.
\end{eqnarray*}
Note that
\begin{eqnarray}
I_{1} &=&\int_{-\infty }^{+\infty }\left\{ \int_{m}^{+\infty
}e^{A_{1}+\sigma _{1}\sqrt{T-t}x}p(x,y,\eta )dy\right\} dx \nonumber \\
&=&\int_{m}^{+\infty }\frac{1}{\sqrt{2\pi }}e^{A_{1}-\frac{1}{2(1-\eta ^{2})}%
\left( y^{2}-\left( \eta y+(1-\eta ^{2})\sigma _{1}\sqrt{T-t}\right)
^{2}\right) }dy \nonumber \\
&=&e^{A_{1}+\frac{1}{2}(1-\eta ^{2})\sigma _{1}^{2}(T-t)+\eta ^{2}\sigma
_{1}^{2}(T-t)}\int_{m}^{+\infty }\frac{1}{\sqrt{2\pi }}e^{-\frac{1}{2}\left(
y-\eta \sigma _{1}\sqrt{T-t}\right) ^{2}}dy \nonumber \\
&=&e^{A_{1}+\frac{1}{2}\sigma _{1}^{2}(T-t)}\Phi (\eta \sigma _{1}\sqrt{T-t}%
-m) \label{i1}
\end{eqnarray}
and
\begin{eqnarray}
I_{2} &=&\int_{-\infty }^{+\infty }\left\{ \int_{m}^{+\infty
}e^{A_{2}+\sigma _{1}\sqrt{T-t}x-by}p(x,y,\eta )dy\right\} dx \nonumber \\
&=&\int_{m}^{+\infty }e^{A_{2}-by}\frac{1}{\sqrt{2\pi }}e^{-\frac{1}{%
2(1-\eta ^{2})}\left[ y^{2}-\left( \eta y+(1-\eta ^{2})\sigma _{1}\sqrt{T-t}%
\right) ^{2}\right] }dy \nonumber \\
&=&e^{A_{2}+\frac{1}{2}\sigma _{1}^{2}(T-t)(1-\eta ^{2})+\frac{1}{2}(\eta
\sigma _{1}\sqrt{T-t}-b)^{2}}\int_{m}^{+\infty }\frac{1}{\sqrt{2\pi }}e^{-%
\frac{1}{2}\left( y-\eta \sigma _{1}\sqrt{T-t}+b\right) ^{2}}dy \nonumber \\
&=&e^{A_{2}+\frac{1}{2}\sigma _{1}^{2}(T-t)-\eta b\sigma _{1}\sqrt{T-t}+%
\frac{b^{2}}{2}}\Phi (\eta \sigma _{1}\sqrt{T-t}-m-b) \label{i2}
\end{eqnarray}
where $\Phi (\cdot )$ is the standard Gaussian distribution function.
Therefore, by (\ref{ii}), (\ref{i1}), and (\ref{i2}), we obtain
\begin{eqnarray*}
&&E_{t}^{Q}\left[ e^{-r(T-t)}\left[ S_{1}(T)-S_{2}(T)\right] ^{+}\right] \\
&=&S_{1}(t)\Phi (\eta \sigma _{1}\sqrt{T-t}-m)-S_{2}(t)\Phi (\eta \sigma _{1}%
\sqrt{T-t}-m-b).
\end{eqnarray*}
To sum up, for constant interest rate $r$, the price of a call exchange
option at time $t$, is given by
\begin{equation}
E_{t}^{Q}\left[ e^{-r(T-t)}\left[ S_{1}(T)-S_{2}(T)\right] ^{+}\right]
=S_{1}(t)\Phi (d_{1})-S_{2}(t)\Phi (d_{2}) \label{qq0}
\end{equation}
where
\begin{eqnarray}
d_{1} &=&\frac{\log [S_{1}(t)/S_{2}(t)]}{\sqrt{(\sigma _{1}^{2}+\sigma
_{2}^{2}-2\rho _{12}\sigma _{1}\sigma _{2})(T-t)}}+\frac{1}{2}\sqrt{(\sigma
_{1}^{2}+\sigma _{2}^{2}-2\rho _{12}\sigma _{1}\sigma _{2})(T-t)},
\label{qq1} \\
d_{2} &=&d_{1}-\sqrt{(\sigma _{1}^{2}+\sigma _{2}^{2}-2\rho _{12}\sigma
_{1}\sigma _{2})(T-t)}, \label{qq2}
\end{eqnarray}
and $\Phi (\cdot )$ is the standard Gaussian distribution function. This is
consistent with the result of Margrabe derived with partial differential
equation approach\cite{marg}.
As shown above, the exchange option pricing can also be obtained within the
framework of the risk neutral measure, consistent with the result obtained
by Margrabe, which was derived with partial differential equation approach.
In this case of constant interest rate, we wish to note that the interest
rate $r$ does not enter the pricing formula explicitly. This special feature
motivates us to look into the issue of pricing exchange option when interest
rate is stochastic in a general form. The next subsection discusses this in
details within the framework of risk-neutral measures.
\subsection{Stochastic interest rate}
In this subsection, we discuss the issue of pricing exchange options when
interest rate is stochastic. It will be shown below that the option price
closed form can be found for most general one-factor stochastic interest
rate processes ( i.e. one Wiener process ). Our result applies to the cases
where one describes the interest rate such as Vasicek term structure, CIR
term structure. These are special cases of our consideration.
In the following, it is assumed that we are always working in the
risk-neutral probability space $(\Omega ,\mathcal{F},Q)$. Each process below
is referred to this risk-neutral probability measure $Q$. Now assume that
the short rate $r$ also follows a Markov diffusion process, that is,
\[
dr(t)=\mu (r(t),t)dt+\sigma (r(t),t)dW_{0}(t).
\]
Here, we do not specify a concrete interest rate model. All we need to
assume is that the interest rate is a Markov diffusion process. The interest
rate is correlated with the two underlying assets of the exchange option
being considered. Assume further that the correlation matrix of $%
[dW_{0}(t),dW_{1}(t),dW_{2}(t)]$ is
\[
\left[
\begin{array}{lll}
1 & \rho _{01} & \rho _{02} \\
\rho _{01} & 1 & \rho _{12} \\
\rho _{02} & \rho _{12} & 1
\end{array}
\right]
\]
where $\rho _{01},$ $\rho _{02},$ and $\rho _{12}$ are constants. Using
Cholesky decomposition, we can obtain the above correlation structure by
setting
\begin{equation}
\left\{
\begin{array}{l}
dW_{0}(t)=dB_{0}(t) \\
dW_{1}(t)=\rho _{01}dB_{0}(t)+\sqrt{1-\rho _{01}^{2}}dB_{1}(t) \\
dW_{2}(t)=\rho _{02}dB_{0}(t)+\frac{\rho _{12}-\rho _{01}\rho _{02}}{\sqrt{%
1-\rho _{01}^{2}}}dB_{1}(t)+\sqrt{1-\rho _{02}^{2}-\frac{(\rho _{12}-\rho
_{01}\rho _{02})^{2}}{1-\rho _{01}^{2}}}dB_{2}(t)
\end{array}
\right. \label{e2}
\end{equation}
where $B_{0}(t),$ $B_{1}(t),$ and $B_{2}(t)$ are three independent standard
Brownian motions. For the underlying assets, their prices will follow the
processes below:
\begin{eqnarray*}
\log [S_{1}(T)] &=&\tilde{A}_{1}+\tilde{\sigma}_{1}\int_{t}^{T}dB_{1}(s) \\
\log [S_{2}(T)] &=&\tilde{A}_{2}+\tilde{\sigma}_{2}\left[ \tilde{\rho}%
_{12}\int_{t}^{T}dB_{1}(s)+\sqrt{1-\tilde{\rho}_{12}^{2}}%
\int_{t}^{T}dB_{2}(s)\right]
\end{eqnarray*}
where
\begin{eqnarray}
\tilde{A}_{1} &=&\log [S_{1}(t)]+\int_{t}^{T}r(s)ds-\frac{1}{2}\sigma
_{1}^{2}(T-t)+\sigma _{1}\rho _{01}\int_{t}^{T}dB_{0}(s), \nonumber \\
\tilde{A}_{2} &=&\log [S_{2}(t)]+\int_{t}^{T}r(s)ds-\frac{1}{2}\sigma
_{2}^{2}(T-t)+\sigma _{2}\rho _{02}\int_{t}^{T}dB_{0}(s), \nonumber \\
\tilde{\sigma}_{1} &=&\sigma _{1}\sqrt{1-\rho _{01}^{2}}, \label{ss1} \\
\tilde{\sigma}_{2} &=&\sigma _{2}\sqrt{1-\rho _{02}^{2}}, \label{ss2} \\
\tilde{\rho}_{12} &=&\frac{\rho _{12}-\rho _{01}\rho _{02}}{\sqrt{1-\rho
_{01}^{2}}\sqrt{1-\rho _{02}^{2}}}. \label{ss3}
\end{eqnarray}
And hence, by conditional expectation, we can price a call exchange option
as
\begin{eqnarray}
&&E_{t}^{Q}\left[ e^{-\int_{t}^{T}r(s)ds}\ \left[ S_{1}(T)-S_{2}(T)\right]
^{+}\right] \nonumber \\
&=&E_{t}^{Q}\left[ e^{-\int_{t}^{T}r(s)ds}E_{t}^{Q}\left( \left[
S_{1}(T)-S_{2}(T)\right] ^{+}\left| \{B_{0}(s):t\leq s\leq T\}\right.
\right) \right] . \label{ex1}
\end{eqnarray}
Note that given a sample path of $\{B_{0}(s):t\leq s\leq T\}$, using the
results in (\ref{i1}) and (\ref{i2}), we obtain the following:
\begin{eqnarray}
&&E_{t}^{Q}\left( \left[ S_{1}(T)-S_{2}(T)\right] ^{+}\left|
\{B_{0}(s):t\leq s\leq T\}\right. \right) \label{ex2} \\
&=&e^{\tilde{A}_{1}+\frac{1}{2}\tilde{\sigma}_{1}^{2}(T-t)}\Phi (\tilde{\eta}%
\tilde{\sigma}_{1}\sqrt{T-t}-\tilde{m})-e^{\tilde{A}_{2}+\frac{1}{2}\tilde{%
\sigma}_{1}^{2}(T-t)-\tilde{\eta}\tilde{b}\tilde{\sigma}_{1}\sqrt{T-t}+\frac{%
\tilde{b}^{2}}{2}}\Phi (\tilde{\eta}\tilde{\sigma}_{1}\sqrt{T-t}-\tilde{m}-%
\tilde{b}) \nonumber
\end{eqnarray}
where
\begin{eqnarray*}
\tilde{\eta} &\equiv &\frac{\tilde{\sigma}_{1}-\tilde{\sigma}_{2}\tilde{\rho}%
_{12}}{\sqrt{\tilde{\sigma}_{1}^{2}+\tilde{\sigma}_{2}^{2}-2\tilde{\rho}_{12}%
\tilde{\sigma}_{1}\tilde{\sigma}_{2}}}, \\
\tilde{m} &\equiv &\frac{\tilde{A}_{2}-\tilde{A}_{1}}{\sqrt{(\tilde{\sigma}%
_{1}^{2}+\tilde{\sigma}_{2}^{2}-2\tilde{\rho}_{12}\tilde{\sigma}_{1}\tilde{%
\sigma}_{2})(T-t)}}, \\
\tilde{b} &\equiv &\sqrt{(\tilde{\sigma}_{1}^{2}+\tilde{\sigma}_{2}^{2}-2%
\tilde{\rho}_{12}\tilde{\sigma}_{1}\tilde{\sigma}_{2})(T-t)}.
\end{eqnarray*}
And it is straightforward to check
\begin{eqnarray}
\tilde{\eta}\tilde{\sigma}_{1}\sqrt{T-t}-\tilde{m} &=&\frac{\log
[S_{1}(t)/S_{2}(t)]+\frac{T-t}{2}(\sigma _{2}^{2}\rho _{02}^{2}-\sigma
_{1}^{2}\rho _{01}^{2})+(\sigma _{1}\rho _{01}-\sigma _{2}\rho
_{02})\int_{t}^{T}dB_{0}(s)}{\sqrt{(\tilde{\sigma}_{1}^{2}+\tilde{\sigma}%
_{2}^{2}-2\tilde{\rho}_{12}\tilde{\sigma}_{1}\tilde{\sigma}_{2})(T-t)}}
\nonumber \\
&&+\frac{1}{2}\sqrt{(\tilde{\sigma}_{1}^{2}+\tilde{\sigma}_{2}^{2}-2\tilde{%
\rho}_{12}\tilde{\sigma}_{1}\tilde{\sigma}_{2})(T-t)}, \\
\tilde{\sigma}_{1}^{2}+\tilde{\sigma}_{2}^{2}-2\tilde{\rho}_{12}\tilde{\sigma%
}_{1}\tilde{\sigma}_{2} &=&(\sigma _{1}^{2}+\sigma _{2}^{2}-2\rho
_{12}\sigma _{1}\sigma _{2})-(\sigma _{1}\rho _{01}-\sigma _{2}\rho
_{02})^{2}. \label{ex4} \\
e^{\tilde{A}_{1}+\frac{1}{2}\tilde{\sigma}_{1}^{2}(T-t)}
&=&S_{1}(t)e^{\sigma _{1}\rho _{01}x-\frac{1}{2}\sigma _{1}^{2}\rho
_{01}^{2}(T-t)}, \label{ex5} \\
e^{\tilde{A}_{2}+\frac{1}{2}\tilde{\sigma}_{1}^{2}(T-t)-\tilde{\eta}\tilde{b}%
\tilde{\sigma}_{1}\sqrt{T-t}+\frac{\tilde{b}^{2}}{2}} &=&S_{2}(t)e^{\sigma
_{2}\rho _{02}x-\frac{1}{2}\sigma _{2}^{2}\rho _{02}^{2}(T-t)}. \label{ex6}
\end{eqnarray}
Since $\int_{t}^{T}dB_{0}(s)\left| \mathcal{F}_{t}\right. \stackrel{D}{%
\thicksim }\mathcal{N}\left( 0,T-t\right) ,$ by (\ref{ex1}), $\cdots ,$ (\ref
{ex6}), we then obtain exchange option price with stochastic interest rates:
Under the conditions $(C1)$ the prices of two underlying assets follow the
geometric Brownian motions (\ref{ee5}), and there is no dividend paying
during the option's life; $(C2)$ If the risk-free rate $r$ is stochastic$,$
then the price of a call exchange option at time $t$, defined by (\ref{ee6}%
), is given by
\begin{eqnarray*}
&&E_{t}^{Q}\left[ e^{-\int_{t}^{T}r(s)ds}\ \left[ S_{1}(T)-S_{2}(T)\right]
^{+}\right] \\
&=&S_{1}(t)\int_{-\infty }^{+\infty }\ \phi (x;\sigma _{1}\rho
_{01},T-t)\Phi (d_{1}(x))dx-S_{2}(t)\int_{-\infty }^{+\infty }\phi (x;\sigma
_{2}\rho _{02},T-t)\Phi (d_{2}(x))\ dx
\end{eqnarray*}
where
\begin{eqnarray*}
d_{1}(x) &=&\frac{\log \left[ S_{1}(t)/S_{2}(t)\right] +\frac{T-t}{2}(\sigma
_{2}^{2}\rho _{02}^{2}-\sigma _{1}^{2}\rho _{01}^{2})+(\sigma _{1}\rho
_{01}-\sigma _{2}\rho _{02})x}{\sqrt{\left[ (\sigma _{1}^{2}+\sigma
_{2}^{2}-2\rho _{12}\sigma _{1}\sigma _{2})-(\sigma _{1}\rho _{01}-\sigma
_{2}\rho _{02})^{2}\right] (T-t)}} \\
&&+\frac{1}{2}\sqrt{\left[ (\sigma _{1}^{2}+\sigma _{2}^{2}-2\rho
_{12}\sigma _{1}\sigma _{2})-(\sigma _{1}\rho _{01}-\sigma _{2}\rho
_{02})^{2}\right] (T-t)}, \\
d_{2}(x) &=&d_{1}(x)-\frac{1}{2}\sqrt{\left[ (\sigma _{1}^{2}+\sigma
_{2}^{2}-2\rho _{12}\sigma _{1}\sigma _{2})-(\sigma _{1}\rho _{01}-\sigma
_{2}\rho _{02})^{2}\right] (T-t)},
\end{eqnarray*}
$\Phi (\cdot )$ is the standard Gaussian distribution function and $\phi
(\cdot ;\mu ,\nu )$ denote a Gaussian density function with mean $\mu $ and
variance $\nu $.
Clearly, if $\sigma _{1}\rho _{01}=\sigma _{2}\rho _{02}$ (that is, the
covariance between processes $dr(t)$ and $dS_{1}(t)$ is the same as the
covariance between processes $dr(t)$ and $dS_{2}(t)$), then the pricing
formulae for exchange options are the same for both the stochastic and
deterministic term structures. Therefore, it is attempting to argue that we
could use this solution as a control variate if one wants to do Monte-Carlo
simulation to price spread option with nonzero strike price $K$.
\subsection{When underlying assets pay dividends}
The pricing formula for exchange options above when interest rates are
stochastic is derived with the assumption that the two underlying assets,
such as stocks, pay no dividends during the options' life. However, in case
of the underlying assets also pay constant or known dividends, the question
will become the general spread options pricing problem.
Assume the amount of dividends $d_{1}(t_{1}),$ $\cdots ,$ $d_{1}(t_{m})$ for
the first asset and $d_{2}(s_{1}),$ $\cdots ,$ $d_{2}(s_{k})$ for the second
asset to be paid at the the dates $0<t_{1}<\cdots <t_{m}<T$ and $%
0<s_{1}<\cdots <s_{k}<T$ respectively are known in advance. Without loss of
generality, we can write the dividend streams for both assets by
\[
d_{i}(t_{1}),\cdots ,d_{i}(t_{n}),\qquad i=1,2
\]
where $n$ is the number of dividends payment dates for both assets.
Therefore, at time $t,$ the present values of all future dividends will be
\[
\sum_{j=1}^{n}d_{i}(t_{j})e^{-\int_{t}^{t_{j}}r(\tau )d\tau
}I_{[t,T]}(t_{j}),\qquad i=1,2,
\]
and the values of all dividends paid after time $t$ and compounded at the
risk-free rate till the option's maturity date $T$ is given by
\[
\sum_{j=1}^{n}d_{i}(t_{j})e^{\int_{t_{j}}^{T}r(\tau )d\tau
}I_{[t,T]}(t_{j}),\qquad i=1,2.
\]
By the same argument of Heath and Jarrow (1988), we decompose the capital
gain processes $G_{i}(t)$ into the asset price process $S_{i}(t)$ and the
dividends streams. Assume further that the $G_{i}(t)$ also follows the
geometric Brownian motions, that is,
\[
\frac{dG_{i}(t)}{G_{i}(t)}=\mu _{i}dt+\sigma _{i}dW_{i}(t),\qquad i=1,\text{
}2.
\]
Then the capital gain $G_{i}(t)$ may be written by
\begin{eqnarray*}
G_{i}(t) &=&S_{i}(t)+\sum_{j=1}^{n}d_{i}(t_{j})e^{\int_{t_{j}}^{T}r(\tau
)d\tau }I_{[t_{j},T]}(t) \\
&=&S_{i}(t)+D_{i}(t)
\end{eqnarray*}
where
\[
D_{i}(t)=\sum_{j=1}^{n}d_{i}(t_{j})e^{\int_{t_{j}}^{T}r(\tau )d\tau
}I_{[t_{j},T]}(t).
\]
Since the dynamics of the capital gains processes $G_{i}(t),$ $i=1,$ $2,$
under the martingale measure $Q,$ is
\[
\frac{dG_{i}(t)}{G_{i}(t)}=r(t)dt+\sigma _{i}dW_{i}(t),
\]
and $G_{i}(0)=S_{i}(0)$ and $G_{i}(T)=S_{i}(T)+D_{i}(T),$ $i=1,$ $2.$
Therefore, the risk-neural pricing formula will be
\begin{eqnarray*}
&&E_{t}^{Q}\left[ e^{-\int_{t}^{T}r(s)ds}\ \left[ S_{1}(T)-S_{2}(T)\right]
^{+}\right] \\
&=&E_{t}^{Q}\left[ e^{-\int_{t}^{T}r(s)ds}\ \left[
G_{1}(T)-G_{2}(T)-(D_{1}(T)-D_{2}(T))\right] ^{+}\right] .
\end{eqnarray*}
In this case, we will have to employ numerical method to value the option
price.
\section{Conclusion}
In this work, we have discussed the issue of pricing exchange options and
spread options. Closed form for the exchange option price is provided
explicitly when the interest rate is stochastic. Our result is valid for
most general term structure model of one factor, which includes Vasicek
model, CIR model, and well-known models as special cases. Our result
indicates that only the correlation coefficients between the interest rate
and the underlying assets will affect the exchange option price. In one
special case, a completely explicit form of the option pricing can be
obtained. We have argued that it is possible to use this solution as a
control variate when doing Monte-Carlo simulation to price spread options
for nonzero strike price $K$ and stochastic interest rate.
\section*{Acknowledgment}
We are indebted to Prof. Phelim P. Boyle, Prof. Don McLeish,
Prof. James Redekope, Prof. K. S. Tan, Dr. Zejiang
Yang, Dr. Ti Wang, Dr. Wei Qian, Dr. H. Huang, Dr. W. H. Zou,
Prof. Kevin Wang, Prof. Ramzi Khuri
for discusions and interactions. Any errors
of this article are solely due to ourselves.
\newpage
|
\section*{Acknowledgement}
We are grateful to K. Imura, Y. Morita, and M. Ogata for useful discussion.
This work is supported by Grant-in-Aid for Scientific Research (C) 10640301
from the Ministry of Education, Science, Sports and Culture.
|
\section{Introduction}\label{sec:intro}
Spin-gap phases of one-dimensional (1D) electron systems have been
studied for long times. This research has been motivated by the phase
transitions in 1D organic conductors. In the past decade, the discovery
of high-$T_{\rm c}$ superconductivity strongly stimulated this study.
The spin-gap transition in 1D lattice models had been mainly analyzed by
two approaches: The one is the weak coupling theory based on the
bosonization theory and renormalization group. The other is numerical
calculation in finite-size systems which is free from approximation. In
the former scheme, the existence of the gap is argued by investigating
the backward scattering effect on the fixed point, but the validity of
the result is ensured only in the weak coupling limit. On the other
hand, in numerical calculation, the analysis is done by a direct
evaluation of the gap and the finite-size scaling method. In this
approach, a singular behavior of the gap near the critical point makes
it difficult to make out the instability.
In order to illustrate the difficulty in the determination of the phase
boundary, let us consider the Hubbard model,
\begin{equation}
{\cal H}_{\rm HM}=
-t\sum_{is}(c^{\dag}_{is} c_{i+1,s}+\mbox{H.c})
+U\sum_i n_{i\uparrow}n_{i\downarrow}.
\end{equation}
This model has a spin gap for $U<0$. According to the Bethe-ansatz
result for the charge gap at half-filling\cite{Ovchinnikov} combining a
canonical transformation \cite{Shiba}, we can obtain the asymptotic
behavior of the spin gap near the critical point $U=0$ as
\begin{equation}
\Delta E\sim \sqrt{2t|U|}{\rm e}^{-\pi t/|U|}.
\label{eqn:gap_Hubbard}
\end{equation}
Since the gap opens slowly near the critical point, it is very difficult
to find the critical point using conventional finite-size scaling
method.
In this paper, we give a remedy for this
problem\cite{Nakamura-N-K,Nakamura}. The many-body problem is often
simplified by using the notion of universality. Generally, 1D electron
systems belong to the universality class of Tomonaga-Luttinger (TL)
liquids\cite{Tomonaga-L,Haldane,Emery,Solyom,Voit} which are
characterized by gapless charge and spin excitations and power-law decay
of correlation functions. This behavior can be described by the
bosonization theory or the $c=1$ conformal field theory (CFT). In this
scheme, the phase transition to the spin-gap phase is understood as an
instability caused by the backward scattering process using the
renormalization group technique\cite{Manyhard-S}, and a spin gap opens
when the backward scattering turns from repulsive to attractive. This
transition point is equivalent to the level-crossing of the
singlet-triplet excitation
spectra\cite{Affleck-G-S-Z,Ziman-S,Okamoto-N}, by taking account of the
logarithmic corrections originated from the backward scattering.
In this paper, we will analyze the following models based on this
notion. The first example is the extended Hubbard model which is
given by
\begin{equation}
{\cal H}_{\rm EHM}={\cal H}_{\rm HM}+V\sum_i n_{i}n_{i+1}.
\label{eqn:EHM}
\end{equation}
For the study of spin-gap transitions, this model has been analyzed by
the {\it g-ology} for weak coupling region\cite{Emery,Solyom}. The
numerical calculation was performed by the exact diagonalization with
finite-size scaling method\cite{Penc-M,Sano-O,Lin-G-C-F-G}. However,
the spin-gap phase boundary has not been clarified.
The next example is the $t$-$J$ model described by
\begin{eqnarray}
{\cal H}_{t\mbox{-}J}&=&
-t\sum_{is}(\tilde{c}^{\dag}_{is} \tilde{c}_{i+1,s}+\mbox{H.c.})\nonumber\\
&&+J\sum_{i}(\bm{S}_i\cdot\bm{S}_{i+1}-n_i n_{i+1}/4),
\label{eqn:t-J}
\end{eqnarray}
where $\tilde{c}_{is}=c_{is}(1-n_{i,-s})$. This model is obtained by
doping holes in the Heisenberg spin chain. For this model, the weak
coupling treatment is difficult due to this strong coupling constraint,
however, the universality class of this model is known as TL liquids,
from the analysis for the exactly solvable cases at $J/t=0$ (spinless
fermion) and $J/t=2$ (super-symmetric
point)\cite{Bares-B-O,Kawakami-Y90b}. The remaining region was analyzed
using the exact diagonalization by Ogata {\it et al.}\cite{Ogata-L-S-A}.
Their phase diagram shows the enhancement of the superconducting
correlation ($K_{\rho} > 1$) and the phase separation
($K_{\rho}\rightarrow\infty$) for the large $J/t$ region. According to
their result, the spin-gap phase does not exist except for the low
density region. Variational approaches also played roles
in the analysis for this model\cite{Hellberg-M,Chen-L,Yokoyama-O}, but
could not establish clear solution for the spin-gap phase.
Extensions of the $t$-$J$ model are also considered by many researchers
\cite{Sano-T,Ogata-L-R,Imada,Dagotto-R,Troyer-T-R-R-D,Ammon-T-T}.
In spin systems, a spin gap opens by the effect of frustration or
dimerization. Metallic spin-gap phases can be generated by doping holes
in these spin systems. In this paper, we concentrate our attention on
the $t$-$J$-$J'$ model which includes the effect of
frustration\cite{Sano-T,Ogata-L-R}:
\begin{eqnarray}
{\cal H}_{t\mbox{-}J\mbox{-}J'}&=&
{\cal H}_{t\mbox{-}J}+J'\sum_{i}(\bm{S}_i\cdot\bm{S}_{i+2}-n_i n_{i+2}/4).
\label{eqn:t-J-J'}
\end{eqnarray}
We introduce a parameter $\alpha$ for the strength of the frustration
given by $\alpha\equiv J'/J$. At half-filling ($n=1$), this model
becomes an $S=1/2$ frustrated spin chain. In this case, the ground state
at $\alpha=1/2$ is the two-fold degenerate dimer state with a spin gap,
and the ground state energy density is
$-3/4J$\cite{Majumder-G,Shastry-S,Broek}. The fluid-dimer transition
occurs at $\alpha_{\rm c}=0.2411$\cite{Okamoto-N}. Upon doping of holes,
the system may become metallic, and the spin gap is reduced\cite{Sano-T}
but persists for the finite doping. The phase diagram of this model for
$n\neq 1$ at $\alpha=1/2$, using the exact diagonalization, was obtained
by Ogata, Luchini, and Rice\cite{Ogata-L-R}, but the phase boundary of
the spin-gap phase was also remained to be ambiguous.
This paper is organized as follows. In Sec.\ref{sec:formalism}, we
discuss, based on the continuum field theory, that the level crossing of
singlet and triplet excitation spectra gives the critical point of the
spin-gap transition. In Sec.\ref{sec:symmetry}, we consider boundary
conditions for the unique ground state, and discrete symmetries of wave
functions to identify the energy spectra observed in our analysis. In
Sec.\ref{sec:diagrams}, we analyze representative models introduced
above, and clarify the spin-gap region in the phase diagrams, and check
the consistency of our argument. Finally, in Sec.\ref{sec:summary}, we
present our conclusions.
The paper also contains three Appendices. The first shows the relation
among the different notations for the quantum numbers. The second is
derivation for the logarithmic corrections. The third explains the
calculation in two-electron systems.
\section{Continuum Field Theory}
\label{sec:formalism}
\subsection{Effective Hamiltonian}
Let us start our argument from the Abelian bosonization theory of
electrons\cite{Haldane,Emery,Solyom,Voit,Schulz}. The low-energy
excitations are described by continuous fermion fields which are defined
by
\begin{equation}
c_{j,s}\rightarrow \psi_{{\rm L},s}(x)+\psi_{{\rm R},s}(x)
\end{equation}
The boson representation of the fermion operator is
\begin{equation}
\psi_{r,s}(x)=\frac{1}{\sqrt{2\pi\alpha}}
{\rm e}^{{\rm i} r k_{\rm F} x}{\rm e}^{{\rm i}/\sqrt{2}\cdot
[r(\phi_{\rho}+s\phi_{\sigma})-\theta_{\rho}-s\theta_{\sigma}]},
\label{eqn:fermion_op}
\end{equation}
where $\alpha$ is a short-distance cutoff. $r={\rm R},{\rm L}$ and
$s=\uparrow, \downarrow$ refer to $+,-$ in that order. The phase fields
are defined as
\begin{mathletters}
\label{eqn:phase_fields}
\begin{eqnarray}
\phi_{\nu}(x)
&=&-\frac{{\rm i}\pi}{L}\sum_{p\neq 0}A_p(x)
\left[\nu_{\rm R}(p) + \nu_{\rm L}(p)\right]
-\frac{\sqrt{2}\pi x}{L}\hat{n}_{\nu},\\
\theta_{\nu}(x)
&=&+\frac{{\rm i}\pi}{L}\sum_{p\neq 0}A_p(x)
\left[\nu_{\rm R}(p) - \nu_{\rm L}(p)\right]
+\frac{\sqrt{2}\pi x}{L}\hat{m}_{\nu},
\end{eqnarray}
\end{mathletters}
where $A_p(x)\equiv\frac{1}{p}{\rm e}^{-{\rm i}\alpha|p|/2-{\rm i}px}$,
and $\nu_{r}$ is the charge ($\nu=\rho$) or the spin ($\nu=\sigma$)
density operator. These phase fields satisfy the relation
$\left[\phi_{\nu}(x),\theta_{\nu}(x')\right]=-{\rm i}\pi
{\rm\,sign}(x-x')/2$.
Using above relations, effective Hamiltonian of a 1D electron
system is described by the U(1) Gaussian model (charge part) and the
SU(2) sine-Gordon model (spin part),
\begin{equation}
{\cal H}={\cal H}_{\rho} + {\cal H}_{\sigma}
+\frac{2 g_{1\perp}}{(2\pi\alpha)^2}
\int_0^L {\rm d}x \cos(\sqrt{8}\phi_{\sigma})
\label{eqn:effHam}.
\end{equation}
Here $g_{1\perp}$ is the backward
scattering amplitude and for $\nu=\rho,\sigma$
\begin{equation}
{\cal H}_{\nu}=\frac{v_{\nu}}{2\pi}\int_0^L {\rm d}x
\left[K_{\nu}(\partial_x \theta_{\nu})^2
+K_{\nu}^{-1}(\partial_x \phi_{\nu})^2\right]
\label{eqn:Gaussian},
\end{equation}
where $v_{\nu}$ and $K_{\nu}$ are the velocity and the Gaussian
coupling, respectively, for the charge ($\nu=\rho$) and the spin
($\nu=\sigma$) sectors. In the TL phase ($g_{1\perp}>0$), the parameters
$K_{\sigma}$ and $g_{1\perp}$ are renormalized as $K^{*}_{\sigma}=1$ and
$g_{1\perp}^{*}=0$, reflecting the SU(2) symmetry.
The phase fields defined in eqs.(\ref{eqn:phase_fields}) satisfy the
following boundary conditions,
\begin{mathletters}
\begin{eqnarray}
\phi_{\nu}(x+L)&=&\phi_{\nu}(x)-\sqrt{2}\pi n_{\nu},\\
\theta_{\nu}(x+L)&=&\theta_{\nu}(x)+\sqrt{2}\pi m_{\nu}.
\end{eqnarray}\label{eqn:BC_phase}
\end{mathletters}
The quantum numbers $m_{\nu}$ and $n_{\nu}$ are defined by the eigen
values of the total number operators $\hat{N}_{r,s}$ (measured with
respect to the ground state) for right and left going particles ($r={\rm
R},{\rm L}$) of spin $s$
\begin{mathletters}
\begin{eqnarray}
n_{\nu}&=&
[(N_{{\rm R}\uparrow}+N_{{\rm L}\uparrow})
\pm (N_{{\rm R}\downarrow}+N_{{\rm L}\downarrow})]/2,\\
m_{\nu}&=&
[(N_{{\rm R}\uparrow}-N_{{\rm L}\uparrow})
\pm (N_{{\rm R}\downarrow}-N_{{\rm L}\downarrow})]/2.
\end{eqnarray}\label{eqn:q_num}
\end{mathletters}
Here the upper and lower sign refer to charge ($\nu=\rho$) and spin
($\nu=\sigma$) degrees of freedoms, respectively. Thus $n_{\nu}$ denotes
excitations involving the variation of particles numbers and $m_{\nu}$
indicate current excitations. If we require $N_{r,s}$ to be an integer,
the possible value of the quantum numbers are restricted as
\begin{equation}
(-1)^{m_{\rho}\pm m_{\sigma}}=(-1)^{n_{\rho}\pm n_{\sigma}}.
\label{eqn:sel_rule}
\end{equation}
This is the selection rule for the quantum
numbers\cite{Haldane,Woynarovich}.
\subsection{Excitation Spectra and Boundary Conditions}
First, we consider the excitation spectra for $g_{1\perp}=0$ case. If
the system is periodic, and has unique ground state, the ground state
energy of the system with length $L$ is given
by\cite{Blote-C-N}
\begin{equation}
E_0(L)=L\epsilon_0-\frac{\pi(v_{\rho}+v_{\sigma})}{6L}c,
\end{equation}
where the central charge $c$ characterizes the universality class of the
model. The finite-size corrections for the excitation energy and
momentum of the system are described by\cite{Cardy84,Haldane}
\begin{eqnarray}
E-E_0&=&
\frac{2\pi v_{\rho}}{L}x_{\rho}+\frac{2\pi v_{\sigma}}{L}x_{\sigma},
\label{eqn:energy}\\
P-P_0&=&\frac{2\pi}{L}(s_{\rho}+s_{\sigma})+2m_{\rho}k_{\rm F},
\label{eqn:momentum}
\end{eqnarray}
where $k_{\rm F}=\pi N/2L$ is the Fermi wave number. $x_\nu=\Delta_\nu^+
+ \Delta_\nu^-, s_\nu=\Delta_\nu^+ - \Delta_\nu^-$ are the scaling
dimension and the conformal spin, respectively, where the conformal
weights for each sector are given by
\begin{equation}
\Delta_{\nu}^{\pm}=\frac{1}{2}
\left(\sqrt{\frac{K_{\nu}}{2}}m_{\nu}
\pm\frac{n_{\nu}}{\sqrt{2K_{\nu}}}\right)^2+n_{\nu}^{\pm}.
\label{eqn:weight}
\end{equation}
Here the integer $n_{\nu}^{\pm}$ denote descendant fields which describe
particle-hole excitations near the Fermi points. The scaling dimensions
are related to the critical exponents for the correlation functions as
\begin{equation}
\langle{\cal O}_i(r){\cal O}_i(r')\rangle
\sim |r-r'|^{-2(x_{\rho i}+x_{\sigma i})}.
\end{equation}
Therefore, there is one to one correspondence between the excitation
spectra and the operators. The operators correspond to the excited
states are given by
\begin{equation}
{\cal O}_{m_{\rho},m_{\sigma},n_{\rho},n_{\sigma}}
\propto{\rm e}^{{\rm i}\sqrt{2}(m_{\rho}\phi_{\rho}+m_{\sigma}\phi_{\sigma}
+n_{\rho}\theta_{\rho}+n_{\sigma}\theta_{\sigma})},
\end{equation}
or their linear combinations. From eqs.(\ref{eqn:fermion_op}) and
(\ref{eqn:BC_phase}), the Fermi operator takes the following boundary
conditions depending on the excited states:
\begin{equation}
\psi_{r,s}(x+L)=\psi_{r,s}(x)
{\rm e}^{{\rm i}\pi(m_{\rho}+m_{\sigma}+n_{\rho}+n_{\sigma})}.
\label{eqn:BC_fermion}
\end{equation}
This means that the excited states given by arbitrary combination of
quantum numbers are realized by changing the boundary conditions, while,
for fixed boundary conditions, the possible excited states are
restricted by the selection rule (\ref{eqn:sel_rule}).
The excitation spectra on which we will turn our attention can be
obtained based on the operators for the charge-density-wave (CDW) and
the spin-density-wave (SDW):
\begin{mathletters}
\begin{eqnarray}
{\cal O}_{\rm CDW}
&=&\psi^{\dag}_{{\rm L}\uparrow}\psi_{{\rm R}\uparrow}
+\psi^{\dag}_{{\rm L}\downarrow}\psi_{{\rm R}\downarrow}\nonumber\\
&=&\frac{1}{\pi\alpha}
\exp({\rm i}2k_{\rm F}x+{\rm i}\sqrt{2}\phi_{\rho})
\cos(\sqrt{2}\phi_{\sigma}),\\
{\cal O}_{{\rm SDW},z}
&=&\psi^{\dag}_{{\rm L}\uparrow}\psi_{{\rm R}\uparrow}
-\psi^{\dag}_{{\rm L}\downarrow}\psi_{{\rm R}\downarrow}\nonumber\\
&=&\frac{{\rm i}}{\pi\alpha}
\exp({\rm i}2k_{\rm F}x+{\rm i}\sqrt{2}\phi_{\rho})
\sin(\sqrt{2}\phi_{\sigma}),\\
{\cal O}_{{\rm SDW},+}
&=&\psi^{\dag}_{{\rm L}\uparrow}\psi_{{\rm R}\downarrow}\nonumber\\
&=&\frac{1}{2\pi\alpha}
\exp({\rm i}2k_{\rm F}x+{\rm i}\sqrt{2}\phi_{\rho})
\exp(+{\rm i}\sqrt{2}\theta_{\sigma}).
\end{eqnarray}
\label{eqn:operators}
\end{mathletters}
These excitation spectra consist of the charge part which carries the
momentum $2k_{\rm F}$, and the spin part which forms singlet
($\sqrt{2}\cos\sqrt{2}\phi_{\sigma}$) and triplet
($\sqrt{2}\sin\sqrt{2}\phi_{\sigma}, \exp(\pm{\rm
i}\sqrt{2}\theta_{\sigma})$) states. Note that the spin part of the
singlet and the triplet superconducting operators (SS, TS) are obtained
with $k_{\rm F}=0$ and replacing $\phi_\rho\rightarrow\theta_{\rho}$.
If charge-spin separation occurs, the spin excitations in
eqs.(\ref{eqn:operators}) ($m_{\sigma}=1$ or $n_{\sigma}=1$, otherwise
$=0$) can be extracted by using anti-periodic boundary conditions
following eq.(\ref{eqn:BC_fermion}). In the continuum field theory based
on the TL model, the dispersion relation is approximated by linearized
one, so that the deviation from the approximated dispersion become
smaller if the excitation energies become lower by eliminating the
charge excitations. Therefore, the precision of the analysis is enhanced
by twisting the boundary conditions.
The twisted boundary conditions are also important in identification of
excitation spectra. Under anti-periodic boundary conditions, the
momenta of these states are reduced to zero. Then we can define the
parity transformation to classify these spectra. Although the
space-inversion operator and translation operator do not commute, we can
classify these spectra simultaneously by wave numbers and parities, if
the wave number $k$ takes $0$ or $\pi$. From eq.(\ref{eqn:fermion_op}),
the phase fields $\phi_{\nu}$ change under the parity (${\cal P}$:
R$\leftrightarrow$L), and the spin-reversal transformations (${\cal T}$:
$\uparrow\leftrightarrow\downarrow$) as\cite{theta_fields}
\begin{mathletters}
\begin{eqnarray}
{\cal P}:&&\phi_{\sigma}\rightarrow-\phi_{\sigma},\ \ \
\phi_{\rho}\rightarrow-\phi_{\rho}\\
{\cal T}:&&\phi_{\sigma}\rightarrow-\phi_{\sigma}
\end{eqnarray}
\end{mathletters}
Thus operators have discrete symmetries as ${\cal P}={\cal T}=1$ for the
singlet ($\sqrt{2}\cos\sqrt{2}\phi_{\sigma}$), and ${\cal P}={\cal
T}=-1$ for the triplet with $S^z=0$
($\sqrt{2}\sin\sqrt{2}\phi_{\sigma}$). The discrete symmetries of the
wave functions of these excited states are determined by combinations of
those of the ground state and the operators. Further discussion for the
discrete symmetries will be given in the next section.
\subsection{Renormalization Group}
Next, we consider the renormalization ($g_{1\perp}\neq 0$).
By the change of the cut off $\alpha\rightarrow {\rm e}^{{\rm d}l}\alpha$,
the coupling constant $g_{1\perp}$ and $K_\sigma$ are renormalized
as\cite{Kosterlitz}
\begin{mathletters}
\begin{eqnarray}
\frac{{\rm d}y_0(l)}{{\rm d}l}&=&-y_1^{\ 2}(l),\\
\frac{{\rm d}y_1(l)}{{\rm d}l}&=&-y_0(l) y_1(l),
\end{eqnarray}
\label{eqn:RGE}
\end{mathletters}
where $y_0(l)\equiv 2(K_{\sigma}-1), y_1(l)\equiv g_{1\perp}/\pi v_{\sigma}$.
For the SU(2) symmetric case $y_{0}(l)=y_{1}(l)$
(the level-$1$ SU(2) Wess-Zumino-Novikov-Witten (WZNW)
model\cite{Knizhnik-Z,Gepner-W,Affleck-G-S-Z}) and $y_0(l)>0$, the
scaling dimensions of the operators for singlet and triplet excitations
split logarithmically by the marginally irrelevant coupling
as\cite{Giamarchi-S,Affleck-G-S-Z} (see Appendix \ref{apx:derivation})
\begin{mathletters}\label{eqn:sclngdim}
\begin{eqnarray}
x_{\sigma}^{\rm singlet}&=&\frac{1}{2}+\frac{3}{4}\frac{y_0}{y_0\ln L+1},\\
x_{\sigma}^{\rm triplet}&=&\frac{1}{2}-\frac{1}{4}\frac{y_0}{y_0\ln L+1},
\end{eqnarray}
\end{mathletters}
where $y_0\equiv y_0(0)$ and we have set $l=\ln L$. When $y_0<0$,
$y_0(l)$ is renormalized to $y_0(l)\rightarrow -\infty$, then a spin gap
appears. At the critical point ($y_0=0$), there are no logarithmic
corrections in the excitation gaps (the logarithmic correction from
higher order also vanish). Therefore, the critical point is obtained
from the intersection of the singlet and the triplet excitation
spectra\cite{Affleck-G-S-Z,Ziman-S,Okamoto-N}. In this case, we can
determine the critical point with high precision\cite{Okamoto-N}, since
the remaining correction is only $x_{\nu}=4$ irrelevant
fields\cite{Cardy86,Reinicke}. This irrelevant field, which does not
exist in the pure sine-Gordon model, comes from the nonlinear term
neglected when linearizing the dispersion relation near the Fermi level
in the course of the bosonization.
The physical meaning of this transition point ($y_0=0$) is the one where
the backward scattering coupling changes from repulsive to
attractive. Moreover, at the critical point, the SU(2) symmetry is
enhanced to the chiral SU(2)$\times$SU(2) symmetry\cite{Affleck-G-S-Z},
since the spin degrees of freedom of the right and the left Fermi points
become independent.
Eq.(\ref{eqn:sclngdim}) also explains the fact that the SDW (CDW)
correlation is dominant for $K_{\rho}<1$ region with(out) spin gap,
while for $K_{\rho}>1$, the TS (SS) correlation is dominant with(out)
spin gap\cite{Giamarchi-S}.
Finally, let us consider the massive region. The behavior of the gap is
explained from the two-loop renormalization group equation of the
level-1 SU(2) WZNW model\cite{Amit-G-G,Destri,Nomura}
\begin{equation}
\frac{{\rm d}y_0(l)}{{\rm d}l}=-y_0^{\ 2}(l)-\frac{1}{2}y_0^{\ 3}(l).
\end{equation}
If we define the correlation length $\xi$ as $y_0(\ln\xi)\equiv-1$ and
the energy gap as $\Delta E=v_{\sigma}/\xi$, then one can derive the
asymptotic form of the spin gap by solving the differential equation for
$|y_0(l)|\ll 1$ as
\begin{equation}
\Delta E\propto\sqrt{|y_{0}|}\exp(-\mbox{Const.}/|y_0|).
\label{eqn:gap}
\end{equation}
Note that eq.(\ref{eqn:gap}) is the same asymptotic behavior as the spin
gap of the negative-$U$ Hubbard model at half-filling given by
eq.(\ref{eqn:gap_Hubbard}).
\section{Uniqueness of Ground State and Discrete Symmetries}
\label{sec:symmetry}
In the previous section, the ground state is assumed to be a singlet, so
that we should consider the way to make the singlet ground state in the
finite-size systems. Furthermore, we also discuss the discrete
symmetries of wave functions to identify the energy
spectra\cite{technical_issue}. The symmetries depend on the choice of
representations for wave functions, so that we consider in the
representative two cases: one is the standard electron systems such as
the (extended) Hubbard model. The other is doped spin chains like the
$t$-$J$ model. In the following argument, the electron hopping is
restricted to the nearest neighbor, and the number of electrons is
assumed to be even. The results are summarized in
Table \ref{fig:symmetries}.
\subsection{Hubbard-type Models}
It is convenient to use the following representation of the basis to
describe the Hubbard-type models which permits double occupancy:
\begin{eqnarray}
|\Psi_{\rm A}\rangle&\equiv&\sum_{n_1<\cdots<n_M; n_{M+1}<\cdots<n_N}
\hspace{-1cm}f_{\rm A}(n_1,\cdots,n_M;n_{M+1},\cdots,n_N)
\nonumber\\
&&\times
\prod_{i=1}^M c^{\dag}_{n_i\downarrow}
\prod_{j=M+1}^N c^{\dag}_{n_j\uparrow}|{\rm vac}\rangle,
\label{eqn:basis_A}
\end{eqnarray}
where $1\leq n_i\leq L$ and the periodicity $n_{i}+L\rightarrow n_i$ is
assumed. $N$ is number of electrons, and $M$ is number of electrons with
down spins. We call this representation as ``basis A''. In the case of
the (extended) Hubbard model, all off-diagonal matrix elements of the
Hamiltonian arise from the hopping term. Then the sign of these elements
are negative as far as no electron hops across the boundary. When an
electron moves across the boundary, the sign of the hopping amplitude
changes depending on $M$ reflecting anti-commutation relations of the
Fermi operators. If the periodic boundary conditions are assumed, the
hopping amplitude at the boundary become $+t$ for $M=$ even case, and
$-t$ for $M=$ odd case.
According to the Lieb-Schultz-Mattis theorem\cite{Lieb-S-M} (the
Perron-Frobenius theorem), if all off-diagonal elements of a real
symmetric irreducible matrix are non-positive, then the all vector
elements for the lowest eigen value have the same sign. Therefore if the
lowest eigen state is degenerate, these states can not be
orthogonal. Thus the ground state is proved to be a singlet. In order to
realize this situation in the (extended) Hubbard model with basis A, we
choose anti-periodic boundary conditions for $M=$ even case, and
periodic boundary conditions for $M=$ odd case. Then all off-diagonal
matrix elements become non-positive and the ground state is proved to be
a singlet. This selection of the boundary conditions are equivalent to
those derived from the Bethe-ansatz result of the Hubbard model in the
strong-coupling limit\cite{Ogata-S}.
The $t$-$J$ model (\ref{eqn:t-J}) can also be described by using the
basis A. This Hamiltonian has off-diagonal matrix elements originated
from the exchange interaction, in addition to the hopping term. The
exchange process gives $-J/2$, where the negative sign arises from the
anti-commutation relations of fermions. Thus the all off-diagonal matrix
elements are non-positive if the boundary conditions are chosen as
discussed above. This situation does not change if the three-site term
is added. Therefore the ground state of the $t$-$J$ model is also proved
to be a singlet.
Now we define an operation for the parity transformation (space
inversion) as
\begin{eqnarray}
\lefteqn{{\cal P}f_{\rm A}(n_1,\cdots,n_M;n_{M+1},\cdots,n_N)=}\nonumber\\
&& f_{\rm A}(\bar{n}_M,\cdots,\bar{n}_1;\bar{n}_N,\cdots,\bar{n}_{M+1}),
\label{eqn:def_P_A}
\end{eqnarray}
where $\bar{n}_j\equiv L+1-n_j$. The spin-reversal transformation can be
defined only for $M=N/2$ cases as
\begin{eqnarray}
\lefteqn{{\cal T}f_{\rm A}(n_1,\cdots,n_M;n_{M+1},\cdots,n_N)=}\nonumber\\
&& f_{\rm A}(n_{M+1},\cdots,n_{N};n_1,\cdots,n_{M}).
\label{eqn:def_T_A}
\end{eqnarray}
The eigen values of the operator ${\cal P}$ and ${\cal T}$ can take $\pm
1$. Since the all vector elements of the ground-state wave function have
the same sign, the ground state discussed above satisfies ${\cal
P}={\cal T}= 1$ and $k=0$.
The discrete symmetries of wave functions for the excited states are
determined by combinations of those of the ground state and the
operators given by the bosonization argument. Thus ${\cal P}={\cal
T}=1$ for the singlet ($\sqrt{2}\cos\sqrt{2}\phi_{\sigma}$), and ${\cal
P}={\cal T}=-1$ for the triplet with $S^z=0$
($\sqrt{2}\sin\sqrt{2}\phi_{\sigma}$). On the contrary, for the triplet
state with $S^z=\pm 1$ ($\exp(\pm{\rm i}\sqrt{2}\theta_{\sigma})$), the
symmetries are the same as those of the ground state (${\cal P}=1$,
$k=0$), because the all off-diagonal matrix elements are
non-positive. In this case, the SU(2) symmetry seems to be broken, this
discrepancy of the parity in the triplet states is due to the definition
of eq.(\ref{eqn:def_P_A}) which does not imply the change of the sign
stems from the permutation of fermions in the ${\cal P},{\cal T}$
transformations. If we take account of the anti-commutation relation of
the Fermi operator in these transformations, then we get ${\cal
P}'=(-1)^{M}{\cal P}$ and ${\cal T}'=(-1)^{N/2}{\cal T}$. Thus the SU(2)
symmetry is recovered.
\subsection{Doped Spin Chains}
On the other hand, the models with a constraint to eliminate doubly
occupied sites, such as the $t$-$J$(-$J'$) model, are obtained by doping
holes into $S=1/2$ spin chains. These models are described by the
``basis B'' which is defined by
\begin{equation}
|\Psi_{\rm B}\rangle\equiv\sum_{n_1<\cdots<n_N}
f_{\rm B}(n_1, s_1;\cdots;n_N, s_N)
\prod_{i=1}^N c^{\dag}_{n_i s_i}|{\rm vac}\rangle.
\label{eqn:basis_B}
\end{equation}
Hereafter, we argue based on the $t$-$J$ model. In order to make the
off-diagonal matrix elements stem from the exchange process
non-positive, we introduce a new basis $f'$ with a sign factor as
\begin{equation}
f_{\rm B}(\cdots)\equiv(-1)^{\sum_{j=1}^N j(s_j+1/2)}f_{\rm B}'(\cdots),
\label{eqn:redef_basis_B}
\end{equation}
where $s_j=\pm 1/2$. Note that $j$ denotes the coordinate of the
squeezed spin system\cite{Ogata-S}. If an electron moves across the
boundary, this sign factor changes as
\begin{equation}
(-1)^{\sum_{j=1}^N j(s_j+1/2)}
\rightarrow
(-1)^{\sum_{j=1}^N j(s_j+1/2)}(-1)^M.
\label{eqn:negative_sign}
\end{equation}
Therefore, an additional negative sign appears if $M=$ odd. In addition
to this, a negative sign is also added, reflecting anti-commutation
relations of Fermi operators, so that the all off-diagonal matrix
elements become non-positive if we chose anti-periodic boundary
conditions for $M={\rm even}$, and periodic boundary conditions for
$M={\rm odd}$. Then all $f_{\rm B}'$ have the same sign so that the
ground state is proved to be a singlet. In this condition, the sign of
$f_{\rm B}$ does not change by the shift operation ($n_j\rightarrow
n_j+1$), so that the wave number of the ground state is $k=0$ for any
$M$.
Next, we define parity and spin-reversal transformations as follows:
\begin{eqnarray}
{\cal P}f_{\rm B}(n_1, s_1;\cdots;n_N, s_N)&=&
f_{\rm B}(\bar{n}_N, \bar{s}_1;\cdots;\bar{n}_1, \bar{s}_N),\\
{\cal T}f_{\rm B}(n_1, s_1;\cdots;n_N, s_N)&=&
f_{\rm B}(n_1, -s_1;\cdots;n_N, -s_N),
\end{eqnarray}
where $\bar{n}_j\equiv L+1-n_j$, $\bar{s}_j\equiv s_{N+1-j}$. As we
have proved , all $f_{\rm B}'$ have the same sign in the ground state,
so that we only consider the variation of the sign factor of
eq.(\ref{eqn:redef_basis_B}) in these transformations. One can easily
show that the parity and the spin-reversal transformations bring
additional negative sign only for $M={\rm odd}$ case. Therefore, the
symmetries of the ground-state wave function are ${\cal P}={\cal T}=1$
for $N/2={\rm even}$, ${\cal P}={\cal T}=-1$ for $N/2={\rm odd}$.
The symmetries of the excited states can also be discussed in the same
way as in the case of the basis A. In the basis B, ${\cal P}={\cal
T}=\pm 1$ for the singlet ($\sqrt{2}\cos\sqrt{2}\phi_{\sigma}$) and
${\cal P}={\cal T}=\mp 1$ for the triplet with $S^z=0$
($\sqrt{2}\sin\sqrt{2}\phi_{\sigma}$), where the upper (lower) sign
denotes $N/2=$ even (odd) case. The triplet states with $S^z=\pm 1$ are
${\cal P}=\mp 1$ ($\exp(\pm{\rm i}\sqrt{2}\theta_{\sigma})$). Note that
SU(2) symmetry is conserved in the parity of triplet excitations due to
the $M$ dependence of the parity.
For half-filling ($N=L$), the 1D $t$-$J$ model should be equivalent to
the $S=1/2$ Heisenberg spin chain. The discrete symmetries known for
the Heisenberg chain are ${\cal P}={\cal T}=1$, $k=0$ for $L/2=$ even
and ${\cal P}={\cal T}=-1$, $k=\pi$ for $L/2=$ odd. In this case, the
negative sign in eq.(\ref{eqn:negative_sign}) for $M=$ odd case can be
canceled by wave number $k=\pi$, instead of changing the boundary
conditions. Therefore the symmetries in these two cases are consistent.
Unfortunately, the above proof can not be applied to the $t$-$J$-$J'$
model which includes the anti-ferromagnetic next-nearest-neighbor
interactions. However, boundary conditions and discrete symmetries of
this model are expected to be the same as those of the $t$-$J$ model, as
far as no instability takes place.
\section{Phase Diagrams of The Lattice Models}
\label{sec:diagrams}
Now we start our analysis for the models introduced in
Sec.\ref{sec:intro}. Besides the spin-gap instability, the charge
degrees of freedom is described by the single parameter $K_{\rho}$. For
$K_{\rho}>1$, the superconducting correlations dominant, while for
$K_{\rho}\rightarrow\infty$, a phase separation takes place. In order
to determine $K_{\rho}$, we need two independent physical quantities. In
our case, we calculate the compressibility and the Drude weight. In
finite-size systems, the compressibility is given by\cite{Ogata-L-S-A}
\begin{equation}
\kappa=
\frac{L}{N^2}\left(\frac{E_0(L,N+2)+E_0(L,N-2)-2E_0(L,N)}{4}\right)^{-1}.
\label{eqn:kappa_num}
\end{equation}
where $E_0(L,N)$ is the ground state energy of a system with size $L$
and $N$ electrons ($n\equiv N/L$). On the other hand, the Drude weight
is given by the relation\cite{Kohn}
\begin{equation}
D=\frac{L}{2}\left.\frac{\partial^2E_0(\Phi)}{\partial\Phi^2}\right|_{\Phi=0},
\end{equation}
where $\Phi$ is the flux which penetrates the ring. In the continuum
field theory, these two physical quantities are described by the
parameters of TL liquids\cite{Schulz,Kawakami-Y90a,Frahm-K}. The
compressibility is given as excitation $n_{\rho}=\pm 1$ in
eqs.(\ref{eqn:momentum}),(\ref{eqn:weight}):
\begin{equation}
\frac{1}{n^2\kappa}=\frac{\pi}{2}\frac{v_{\rho}}{K_{\rho}},\label{eqn:kappa}
\end{equation}
and the Drude weight is given by the excitation $m_{\rho}=\Phi/\pi$ as
\begin{equation}
D=\frac{K_{\rho}v_{\rho}}{\pi}.
\label{eqn:drude}
\end{equation}
Therefore, $K_{\rho}$ is obtained as $K_{\rho}=\pi\sqrt{Dn^2\kappa/2}$.
In order to obtain scaling dimensions of the spin degrees of freedom,
the spin-wave velocity is calculated by the following relation,
\begin{equation}
v_{\sigma}=\lim_{L\rightarrow\infty}
\frac{E(L,N,S=1,q=2\pi/L)-E_0(L,N)}{2\pi/L}.
\label{eqn:spin_velocity}
\end{equation}
The extrapolation is done by the function $v_{\sigma}(L) =
v_{\sigma}(\infty) + A/L^2 +B/L^4$.
These corrections are explained by the irrelevant fields with
$x_{\nu}=4$.
In the following, we analyze some models. Since there are too many
instabilities in the extended Hubbard model, we consider the $t$-$J$
model first to turn our attention on the spin-gap instability.
\subsection{The $t$-$J$ Model}
Here we analyze the spin-gap phase of the 1D $t$-$J$ model
(\ref{eqn:t-J}) by the above explained method. We diagonalize $L=8$-$30$
systems by the use of the Lanczos and the Householder algorism.
In order to investigate the structure of excitation spectra in detail,
we show in Fig.\ref{fig:sflow} the spectral flow (flux dependence of
energy) of $N/L=4/8$ system at $J/t=2$. In this case, the boundary
conditions are fixed to the ground state, so that the singlet and the
triplet excitation spectra appear at $\Phi=\pi$ which is equivalent to
the twisted boundary conditions. However, the wave number is not $k=0$
but $k=2k_{\rm F}$. This momentum shift is explained by the relation
$k(\Phi)=k(0)+N\Phi/L$\cite{momentum_shift}. The singlet and the triplet
excitation spectra are connected adiabatically from those of the CDW and
the SDW, respectively. If the linearized dispersion relation is exact,
these two spectra move parallelly versus $\Phi$ in this diagram. As
shown in eqs.(\ref{eqn:operators}), the charge degrees of freedoms
contribute the same amount in the CDW and the SDW spectra, so that the
critical point can be obtained by the level crossing of these
spectra. However, in this case, the precision become lower due to the
irrelevant fields, and the identification by the parity become
impossible due to the incommensurate wave number.
Figure \ref{fig:EXAMPLE} shows the singlet and the triplet excitation
spectra ($L=16,n=1/2$) versus $J/t$. The level crossing takes place at
$J/t\sim 2.7$. The size dependence of the critical point is shown in
Fig.\ref{fig:SIZE-QF}. Since the critical point is almost independent of
the system size, the phase diagram can be constructed without
extrapolation. Then we obtain Fig.\ref{fig:phsdgrm}(a).
In contrast to the former results\cite{Ogata-L-S-A,Hellberg-M,Chen-L},
the spin-gap phase spreads extensively toward the high-density
region. The spin-gap and phase-separation boundaries flow together
into the point $J/t\sim 3.5$ as $n\rightarrow 1$. We are not able to
answer whether the spin gap survives in the $n\rightarrow 1$ limit or
not, because the numerical results become unstable in the high density
region where the phase boundary is close to the phase-separated state.
In the low-density region, the phase boundary can be determined
analytically by solving a two-electron problem (see Appendix
\ref{apx:twoele}). Then the asymptotic behavior of the phase boundary
is obtained as
\begin{equation}
2t/J_{\rm c}=\cos(\pi n/2).
\label{eqn:J_c-in-low-density}
\end{equation}
Note that the $J_{\rm c}$ given by eq.(\ref{eqn:J_c-in-low-density}) in
$L\rightarrow\infty$ limit is equivalent to the critical point where the
singlet pair forms a bound state in the ground state\cite{Lin}. This
explains the fact that the spin-gap phase boundary overlaps the
$K_{\rho}=1$ line where the TL liquid behaves as free electrons, in the
low-density limit.
In order to check the consistency of our argument, we calculate the
scaling dimensions for the singlet and the triplet excitations from
eqs.(\ref{eqn:energy}) and (\ref{eqn:spin_velocity}).
Then the average of the renormalized scaling dimension
(\ref{eqn:sclngdim}) is taken so as to eliminate the logarithmic
corrections as
\begin{equation}
\bar{x}_{\sigma}\equiv
\frac{x_{\sigma}^{\rm singlet}+3x_{\sigma}^{\rm triplet}}{4}.
\label{eqn:average}
\end{equation}
$\bar{x}_{\sigma}$ and its finite-size effect are shown in
Fig.\ref{fig:RATIO} and Fig.\ref{fig:size_xs}, respectively. The
extrapolated data become $1/2$ with error less than 0.2 \%.
In the spin-gap region ($J>J_{\rm c}$), the asymptotic behavior of the
spin gap is obtained using the relation $y_0\propto J-J_{\rm c}$ and
eq.(\ref{eqn:gap}).
\subsection{The $t$-$J$-$J'$ Model}
Next, we analyze the 1D $t$-$J$-$J'$ model (\ref{eqn:t-J-J'}).
According to Ref.\ref{Ogata-L-R}, the critical point for
the $J,J'\rightarrow 0$ limit is obtained by mapping the spin part onto
the case of $n=1$, using the factorized wave function,
\begin{mathletters}
\begin{eqnarray}
J_{\rm eff}&=&J\langle n_i n_{i+1}\rangle_{\rm SF}+
J'\langle n_i (1-n_{i+1})n_{i+2}\rangle_{\rm SF},\\
J'_{\rm eff}&=&J'\langle n_i n_{i+1}n_{i+2}\rangle_{\rm SF},
\end{eqnarray}
\end{mathletters}
where $\langle\cdots\rangle_{\rm SF}$ indicates the expectation value of the
non-interacting spinless fermion. The effective ratio of the frustration
$\alpha_{\rm eff}$ is then obtained as
\begin{equation}
\alpha_{\rm eff}(n,\alpha)=\left[
\frac{(1+1/\alpha)n^{\,2}-s_2^{\,2}-s_1^{\,2}/\alpha}
{n^3-(2s_1^{\,2}+s_2^{\,2})n+2s_1^{\,2}s_2}-1\right]^{-1},
\label{eqn:eff_alpha}
\end{equation}
where $s_l\equiv \sin(l\pi n)/l\pi$. We can obtain the critical density
$n_{\rm c}$ where the spin gap vanishes, by comparing
eq.(\ref{eqn:eff_alpha}) with the result of the frustrated spin chain:
$\alpha_{\rm c}=0.2411$\cite{Okamoto-N}. For $\alpha=1/2$, we get
$n_{\rm c}=0.7433$.
On the other hand, in the low-density limit, the critical value for
the spin-gap phase $J_{\rm c}/t$, can be analytically obtained by
solving the two-electron problem as (see Appendix \ref{apx:twoele})
\begin{equation}
4t/J_{\rm c}=1+2\alpha+\sqrt{1+4\alpha^2}.
\label{eqn:two_elec}
\end{equation}
The meaning of this point is same as that of the $t$-$J$ model.
We show the phase diagram of $\alpha=\alpha_{\rm c}$ case in
Fig.\ref{fig:phsdgrm}(b). The spin-gap phase boundary overlaps with the
contour line of $K_{\rho}=1$ at almost all densities. This situation is
quite resemble to that of the super-symmetric $t$-$J$ model with
long-range hopping and interactions\cite{Kuramoto-Y}. In this model,
there are no logarithmic corrections and the exact ground state is given
by the Gutzwiller wave function. This means that the charge degrees of
freedom is free electrons ($K_{\rho}=1$), and the singlet and the
triplet excitation spectra are degenerate for all densities.
Figure \ref{fig:phsdgrm}(c) is the phase diagram at $\alpha=1/2$. The
phase boundary starts from the critical value of the low-density limit
(\ref{eqn:two_elec}), and bends at $n\sim2/3$. It then flows into the
critical point $(J/t,n)=(0,n_{\rm c})$. Thus the spin-gap phase with
different origin (the Majumder-Ghosh-like dimer phase in the low-doping
region, and the spin-gap phase in the large $J/t$ region) have a single
domain in the phase diagram. In contrast to the case of the $t$-$J$
model ($\alpha=0$), the spin-gap phase boundary lies in the $K_{\rho}<1$
region so that there is no TS region in this case.
Spin-gap phase may also exists for $\alpha>1/2$ cases. For example,
in $\alpha=\infty$ case, $J'_{\rm c}/t=1$ in the dilute limit, and
$n_{\rm c}=0.5752$ in the $J'\rightarrow 0$ limit.
In spite of the deformation of the phase diagram, the critical value
at the quarter-filling ($n=1/2$) is almost independent of the strength
of the frustration $\alpha$, and is kept at $J_{\rm c}/t\sim 2.7$.
Let us consider the reason for this using an argument based on the {\it
g-ology} model\cite{Emery,Solyom}. In order to apply the {\it g-ology}
model, we add the on-site Coulomb term ${\cal H}_U$ to
eq.(\ref{eqn:t-J-J'}) and relax the constraint. The original Hamiltonian
is restored when we set $U=\infty$. Since the {\it g-ology} model is
appropriate for the weak coupling case, we consider $J'$ terms as
corrections to the $t$-$J$ model which belongs to the universality class
of the TL model. Then their contributions to the $g$-parameters, which
are related to the spin-gap generation, are identified as
\begin{equation}
\delta g_{1\perp}=\delta g_{\sigma}
=-J'(1+\cos 4 k_{\rm F}).
\label{eqn:correction}
\end{equation}
For the quarter-filling, eq.(\ref{eqn:correction}) vanishes, so that the
$J'$ terms do not affect the renormalization flow of the spin part. Thus
the frustration does not change the critical point at the
quarter-filling within the scheme of the {\it g-ology} model.
\subsection{The Extended Hubbard Model}
The instability of the extended Hubbard model can be argued based on
the {\it g-ology} model for weak coupling cases ($U,V\rightarrow
0$)\cite{Emery,Solyom}. The $g$ parameters which are used to determine
the phase diagrams are identified as follows:
\begin{mathletters}
\begin{eqnarray}
g_{1\perp}=g_{\sigma}&=&U+2V\cos 2k_{\rm F},\label{eqn:g_sigma}\\
g_{\rho}&=&U+2V(2-\cos 2k_{\rm F}).\label{eqn:g_rho}
\end{eqnarray}
\end{mathletters}
We have defined $g_{\nu}\equiv g_{1\parallel}-g_{2\parallel}\mp
g_{2\perp}$ where the upper (lower) sign corresponds to $\nu=\rho$
($\nu=\sigma$). Then, from the discussion given in
Sec.\ref{sec:formalism}, the spin-gap phase boundary is determined by
$g_{1\perp}=g_{\sigma}=0$. On the other hand, the contour line for
$K_{\rho}=1$ is determined by the condition $g_{\rho}=0$, due to the
following relations
\begin{equation}
K_{\nu}=\sqrt{\frac{2\pi v_{\rm F}+g_{4\parallel}\pm g_{4\perp}+g_{\nu}}
{2\pi v_{\rm F}+g_{4\parallel}\pm g_{4\perp}-g_{\nu}}},
\end{equation}
where $g_{4\perp}=U+2V$, $g_{4\parallel}=2V$.
Figure \ref{fig:tUV} shows the phase diagrams of the 1D extended Hubbard
model (\ref{eqn:EHM}) for various electron fillings. They are obtained
by analyzing the data of $L=12$ systems. In the all cases, the slopes of
the spin-gap phase boundaries and the $K_{\rho}=1$ contour lines near
the origin of the $U$-$V$ plain, are consistent with those predicted by
the {\it g-ology}. For $V<0$ region, there is phase-separated state and
its boundaries flow into $(U,V)=(\infty,-2t)$ due to the equivalence of
the XXZ spin chain in the large-$U$ limit\cite{Schulz}. The spin-gap
phase boundaries flow into these phase-separation boundaries.
In $n=1/3$ case, the spin-gap phase boundary and the $K_{\rho}=1$
contour line almost overlap near the solution of the two electron
problem (see Appendix \ref{apx:twoele}),
\begin{equation}
V_{\rm c}=-\frac{2U_{\rm c}}{U_{\rm c}/t+4}.\label{eqn:two_Vc_tUV}
\end{equation}
This phenomenon is same as that of the low-density region of the
$t$-$J$(-$J'$) model.
At $n=1/2$, the spin-gap phase boundary is close to $U=0$, because the
effect of $V$ is canceled in eq.(\ref{eqn:g_sigma}). In this phase
diagram, there are two regions with $K_{\rho}>1$. Besides of the
spin-gap phase, a charge-gap phase exists for $U,V>0$ region due to the
Umklapp scattering. The analysis for this instability will be reported
elsewhere\cite{Nakamura_98}.
For $n=2/3$, a spin-gap phase appears in $U/t,V/t>0$
region\cite{Sano-O}. This is because the strong nearest-neighbor
repulsion stabilizes the on-cite singlet pairs. The one of the striking
feature in this phase diagram is that there are two phase separated
states in the $V/t\gg 1$ region, and the spin-gap phase boundary flows
between these two phase-separated states. In this region, the spin-gap
phase boundary shifts to the large $U$ side due to the strong
finite-size effect. The phase-separated state in the $U/t>0$ side is
considered as a mixture of $4k_{\rm F}$- and $2k_{\rm F}$-CDW
phases. The stability of this phase is already argued in
Ref.\ref{Sano-O} by using the second-order perturbation theory. On the
other hand, in the $U/t<0$ side, the system is separated into a $2k_{\rm
F}$-CDW phase and a vacuum. These phase-separated states are illustrated
in Fig.\ref{fig:ps_fig}.
The consistency of the argument can also be checked as in the case of
the $t$-$J$ model. Fig.\ref{fig:RATIO_2} shows the averaged scaling
dimension (\ref{eqn:average}) at $n=1/2$ for $V/t=2$ and $8$ cases,
calculated by the data of $L=8,12,16$ systems. Although the finite-size
effect is large for the $V\gg 1$ region, the extrapolated value become
$1/2$.
\section{Conclusion}\label{sec:summary}
To conclude, we have studied critical properties of spin-gap phases in
1D electron systems, considering the effect of the backward scattering
in TL liquids by the renormalization group analysis. The phase boundary
between TL liquids and spin-gap phases is shown to be determined by the
singlet-triplet level crossing point. These excitation spectra are
extracted by twisting boundary conditions, and identified by the
discrete symmetries of wave functions. For this purpose, we have
discussed symmetries of wave functions under parity and spin-reversal
transformations. We have applied the analysis to the extended Hubbard
model and the $t$-$J$-($J'$) model, and clarified the spin-gap regions
in the phase diagrams. The consistency of the our result has been
checked by investigating the ratio of logarithmic corrections. Our
results are also consistent with those of the {\it g-ology} model in the
weak coupling limit, and of the two-electron problem in the dilute
limit.
\section{Acknowledgments}
M. N. thanks E. Dagotto, K. Itoh, K. Kusakabe, M. Ogata, K. Okamoto,
M. Oshikawa, H. Shiba, M. Takahashi, and H. Yokoyama for valuable
comments. A. K. is supported by JSPS Research Fellowships for Young
Scientists. The computation in this work was partly done using the
facilities of the Supercomputer Center, Institute for Solid State
Physics, University of Tokyo.
|
\section{Introduction}
{}In the recent years much progress has been made in understanding six and
four dimensional
orientifold compactifications. Various six dimensional orientifold vacua
were constructed, for instance, in \cite{PS,GP,GJ,bij}. Generalizations of
these constructions to four dimensional orientifold vacua have also been
discussed in detail in \cite{BL,Sagnotti,ZK,KS,Zw,ibanez,KST,blumen,223}.
In many cases the perturbative world-sheet approach to orientifolds gives
rise to consistent anomaly free vacua in six and four dimensions.
However, as was
pointed out in \cite{KST}, there
are cases where the perturbative orientifold description is inadequate as it
misses certain
non-perturbative sectors giving rise to massless states.
In certain cases this inadequacy results in obvious
inconsistencies such as lack of tadpole and anomaly cancellation. Examples
of such cases were discussed in \cite{Zw,ibanez,KST}. In other cases,
however, the issue is more subtle as the non-perturbative states arise in
anomaly free combinations, so that they are easier to miss.
{}Thus, let us consider Abelian $T^6/\Gamma$ orbifold compactifications of
Type I with ${\cal N}=1$ supersymmetry in four dimensions.
The requirement that the orbifold group $\Gamma$ act
crystallographically on $T^6$ restricts the allowed choices of $\Gamma$ to
${\bf Z}_3^*$, ${\bf Z}_4^*$, ${\bf Z}_6^*$, ${\bf Z}_6^\prime$, ${\bf Z}_7$,
${\bf Z}_8$, ${\bf Z}_8^\prime$, ${\bf Z}_{12}$ and ${\bf Z}_{12}^\prime$ for
$\Gamma\approx{\bf Z}_N$, and to ${\bf Z}_2\otimes {\bf Z}_2$,
${\bf Z}_2\otimes {\bf Z}_4$, ${\bf Z}_2\otimes {\bf Z}_6$, ${\bf Z}_2\otimes
{\bf Z}_6^*$, ${\bf Z}_3\otimes
{\bf Z}_3$, ${\bf Z}_3\otimes {\bf Z}_6$ and ${\bf Z}_4\otimes {\bf Z}_4$
for $\Gamma\approx{\bf Z}_N\otimes {\bf Z}_M$. Here we use star and prime
to distinguish different cyclic groups of the same order which act differently
on $T^6$ (the precise actions of these orbifold groups on $T^6$ are given in
the subsequent sections). In some of the above cases
Type I-heterotic duality enables one to argue along the lines of
\cite{ZK,KS,KST} that the non-perturbative states become heavy
and decouple once the
corresponding orbifold singularities are appropriately blown
up. In particular, these are the ${\bf Z}_3^*$ \cite{Sagnotti,ZK},
${\bf Z}_7$, ${\bf Z}_3\otimes {\bf Z}_3$ and ${\bf Z}_6^*$ \cite{KS},
and ${\bf Z}_2\otimes {\bf Z}_6^*$ \cite{223}
orbifold cases. As was argued in \cite{KST}, in other perturbatively tadpole
free orientifolds, such as the ${\bf Z}_6^\prime$,
${\bf Z}_2\otimes{\bf Z}_6$,
${\bf Z}_3\otimes{\bf Z}_6$ and ${\bf Z}_6\otimes{\bf Z}_6$ \cite{Zw}, and
${\bf Z}_{12}$ \cite{ibanez}
orbifold cases, the non-perturbative open string sectors do give rise to
massless states which do not decouple even for blown-up orbifolds.
Recently some examples of non-perturbative orientifolds were explicitly
constructed in \cite{NP}. Thus, the four dimensional example discussed
in \cite{NP} is based on the ${\bf Z}_6^\prime$ orbifold.
The purpose of this paper, which is the follow-up
of \cite{NP}, is to extend the discussions of \cite{NP} to understand
all other orientifold cases as well.
In particular, we will discuss the ${\bf Z}_2\otimes{\bf Z}_6$,
${\bf Z}_3\otimes{\bf Z}_6$, ${\bf Z}_6\otimes{\bf Z}_6$ and
${\bf Z}_{12}$ cases. The first three cases lead to anomaly free
non-perturbative orientifolds. The last case, however, turns out to suffer from
a non-perturbative anomaly whose origins we explain in detail in section V.
Moreover, we elaborate on the rest of the orbifold compactifications mentioned
above, namely, the ${\bf Z}_2\otimes {\bf Z}_4$ and ${\bf Z}_4\otimes
{\bf Z}_4$ \cite{Zw}, and ${\bf Z}_8$, ${\bf Z}_8^\prime$ and
${\bf Z}_{12}$ \cite{ibanez} cases, which
where shown to suffer from perturbative tadpoles in \cite{Zw} and
\cite{ibanez}, respectively. In \cite{KST} this was explained by considering
F-theory duals of these compactifications. In this paper we elaborate the
arguments of \cite{KST}. In particular, we point out that the corresponding
inconsistencies arise due to the particular choices of the gauge bundle
in the models of \cite{Zw,ibanez} (these gauge bundles are perturbative
from the orientifold viewpoint).
{}The origin of non-perturbative states in orientifold compactifications
can already be understood in six dimensions. Thus, in the K3 orbifold examples
of \cite{GJ} the orientifold projection is not $\Omega$, which we will use
to denote that in the smooth K3 case, but rather $\Omega J^\prime$, where
$J^\prime$ maps the $g$ twisted sector to its conjugate $g^{-1}$ twisted
sector (assuming $g^2\not=1$) \cite{pol}. Geometrically this can be viewed
as a permutation of two ${\bf P}^1$'s associated with each fixed point of
the orbifold \cite{KST}. (More precisely, these ${\bf P}^1$'s correspond to
the orbifold blow-ups.) This is different from the orientifold projection
in the smooth case where (after blowing up) the orientifold projection
does {\em not} permute the two ${\bf P}^1$'s. In the case of the
$\Omega J^\prime$ projection the ``twisted'' open string sectors corresponding
to the orientifold elements $\Omega J^\prime g$ are absent
\cite{blum,KST}. However, if the orientifold projection is $\Omega$, then
the ``twisted'' open string sectors corresponding
to the orientifold elements $\Omega g$ are present \cite{KST}.
In fact, these states
are non-perturbative from the orientifold viewpoint and are required for
gravitational anomaly cancellation in six dimensions. In
certain cases Type I-heterotic duality allows one to understand such
sectors and construct the corresponding models explicitly \cite{NP}.
{}In four dimensional orientifolds with ${\cal N}=1$ supersymmetry there
are always sectors (except in the ${\bf Z}_2\otimes {\bf Z}_2$ model of
\cite{BL} which is completely perturbative from the orientifold viewpoint)
such that there is only one ${\bf P}^1$ per fixed point, so only the
$\Omega$ orientifold projection is allowed. This results in
non-perturbative ``twisted'' open string sectors, which, as we have
already mentioned, decouple in certain cases once the appropriate blow-ups
are performed. In other cases we must include these states to obtain the
complete description of a given orientifold.
{}Some of the non-perturbative orientifolds have perturbative heterotic
duals. However, non-perturbative orientifolds with, say, D5-branes are
non-perturbative from the heterotic viewpoint. In this paper we are
therefore exploring some vacua in the region ${\cal D}$ in Fig.1 (which has
been borrowed from the second reference in \cite{223}).
Thus, the ${\bf Z}_6^\prime$ orbifold
compactification discussed in \cite{NP}, as well as the ${\bf Z}_2\otimes
{\bf Z}_6$ orbifold compactification we discuss in section V of this paper,
are examples of four
dimensional {\em chiral} ${\cal N}=1$ supersymmetric string vacua which are
non-perturbative from both orientifold and heterotic points of view.
{}The key point which allows us to understand four dimensional
non-perturbative orientifolds discussed in this paper
is the fact that these models correspond to Type I
compactifications on generalized Voisin-Borcea orbifolds of the form
$(T^2\otimes {\mbox{K3}})/{\bf Z}_N$ ($N=2,3,4,6$). One of the simplifying
features of
these compactifications is that once non-perturbative K3 orientifolds are
understood (and this is partially facilitated by the fact that in six
dimensions the gravitational anomaly cancellation condition is rather
constraining), it is possible to extend the corresponding results to
generalized Voisin-Borcea compactifications. Another nice feature is that,
by T-dualizing along $T^2$, we can map these modes to
F-theory compactifications on the corresponding Calabi-Yau four-folds, and
the (at least partially) geometric picture arising in the F-theory context
is very helpful in understanding such vacua.
{}The rest of this paper is organized as follows. In section II
we use the map between Type I compactifications on K3 and F-theory on
Voisin-Borcea orbifolds to determine which gauge bundles are perturbative from
the orientifold viewpoint. In section III we give explicit models corresponding
to K3 compactifications of Type I with these perturbative gauge bundles.
In section IV we move down to four dimensions and consider Type I
compactifications on Voisin-Borcea orbifolds. In particular, using the map
between such Type I vacua and F-theory on Calabi-Yau four-folds we give the
constraints necessary for tadpole cancellation in the cases where
the gauge bundle is perturbative. In particular, this discussion relies on
the results of section II. In section V we discuss Type I
compactifications on generalized Voisin-Borcea orbifolds
$(T^2\otimes {\mbox{K3}})/{\bf Z}_N$ with $N=3,4,6$. In particular, we argue
that the ${\bf Z}_{12}$ case suffers from a non-perturbative anomaly.
In section VI we discuss some directions for extending our results. In
particular, we consider Type I on K3 with non-zero NS-NS $B$-field and the
gauge bundles with vector structure.
\section{Type I on K3 (Orbifolds)}
{}In this section we discuss six dimensional ${\cal N}=1$ supersymmetric
Type I compactifications on K3. In particular,
what we would like to understand here is which choices of the gauge bundle
are perturbative from the orientifold viewpoint. Our discussion here applies
to both Type I and ${\mbox{Spin}}(32)/{\bf Z}_2$ heterotic compactifications
on K3, which is due to Type I-heterotic duality. The dictionary between the
two descriptions includes mapping 5-branes on the Type I side, which are
made of some number of D5-branes, to NS 5-branes (or, equivalently, small
instantons) on the heterotic side.
{}Thus, consider Type I on a K3 surface. In order to determine a particular
ground state
of the theory, we must specify not only the compactification space,
which in this case is K3, but also the choice of the gauge bundle. Thus,
in the orientifold description of Type I non-perturbative effects can arise
from two different sources: (1) the perturbative orientifold description may
fail to capture all states arising due to, say, various singularities in K3
itself; (2) a given choice of the gauge bundle may not have perturbative
orientifold description. The second source of non-perturbative physics is
generally more non-trivial to understand. Here, however, we will be interested
in determining which choices of the gauge bundle do possess perturbative
orientifold description, and once we understand this class of gauge bundles, we
can then focus on the first source of non-perturbative effects which are easier
to handle due to various hints from geometry.
{}The case of K3 is particularly convenient as (unlike in the case of
Calabi-Yau three-folds) topologically we are dealing with one surface. This
enables us to understand perturbative gauge bundles in the K3 case by going to
orbifold limits which have a simple description. Thus, let us consider orbifold
K3's of the form ${\mbox{K3}}=T^4/{\bf Z}_N$, where $N=2,3,4,6$ are the choices
allowed by the requirement that the orbifold act crystallographically on the
four-torus $T^4$. Let $z_1,z_2$ be the complex coordinates parametrizing
$T^4$. Then the action of the generator $g$ of ${\bf Z}_N$ is given by
\begin{equation}
g z_1=\omega z_1~,~~~gz_2=\omega^{-1} z_2~,
\end{equation}
where $\omega\equiv\exp(2\pi i/N)$.
{}Next, consider Type I on $T^4/{\bf Z}_N$. We will assume that the (untwisted)
NS-NS $B$-field is trivial. Then we can view Type I on $T^4/{\bf Z}_N$ as the
$\Omega$ orientifold of Type IIB on $T^4/{\bf Z}_N$. (More precisely, the
$\Omega$ orientifold is well defined after the appropriate blow-ups of the
orbifold singularities other than ${\bf Z}_2$.) The theory has ${\cal N}=1$
supersymmetry in six dimensions. The massless spectrum of the closed string
sector consists of the gravity supermultiplet, one tensor supermultiplet, and
20 hypermultiplets (which are neutral under the non-Abelian gauge symmetries
in the open string sector). In the open string sector we have 32 D9-branes,
which is required by the untwisted tadpole cancellation conditions. In the
${\bf Z}_3$ case we do not have any D5-branes. In the ${\bf Z}_2,{\bf Z}_4,
{\bf Z}_6$ cases tadpole cancellation conditions also
require introduction of 32 D5-branes.
{}To understand the gauge bundles in all of the above cases, recall that in
the case of K3 anomaly cancellation requires that the number $n_I$ of
instantons embedded in the gauge bundle and the number $n_5$ of 5-branes
(that is, small instantons) must add up to 24: $n_I+n_5=24$.
In the ${\bf Z}_3$ all 24 instantons are embedded in the gauge bundle, so the
number of 5-branes is zero. This gauge bundle is therefore perturbative from
the heterotic viewpoint. On the orientifold side it is defined by the action
of the orbifold group on the Chan-Paton charges which is given by the
following $16\times 16$ matrix\footnote{Throughout this paper
(unless specified otherwise) we work with
$16\times 16$ (rather than $32\times 32$) Chan-Paton matrices for we choose
not to count the orientifold images of the corresponding D-branes.} \cite{GJ}:
\begin{equation}\label{Z_3}
{\bf Z}_3:~~~\gamma_{g}=
{\mbox{diag}}(\alpha {\bf I}_4,\alpha^{-1} {\bf I}_4,{\bf I}_8)~,
\end{equation}
where $\alpha\equiv\exp(2\pi i/3)$, and ${\bf I}_n$ denotes the $n\times n$
identity matrix.
{}Next, consider the ${\bf Z}_2$ case. Here we have 32 D5-branes which form
$n_5=8$
dynamical 5-branes. Here two pairings take place - one due to the orientifold
projection, and the other one due the ${\bf Z}_2$ orbifold projection. We
therefore conclude that $n_I=16$ in this case. This gauge bundle is
non-perturbative from the heterotic viewpoint. On the orientifold side the
corresponding Chan-Paton matrix is given by \cite{PS,GP}:
\begin{equation}\label{Z_2}
{\bf Z}_2:~~~\gamma_{g}={\mbox{diag}}(i{\bf I}_8,-i{\bf I}_8)~.
\end{equation}
{}As to the ${\bf Z}_4$ and ${\bf Z}_6$ cases, the corresponding models
are on the same moduli as the ${\bf Z}_2$ model. Indeed, in all three of
these cases the gauge group can be Higgsed completely. In fact, in the
${\bf Z}_4$ and ${\bf Z}_6$ cases this can already be done just with the
matter arising in the sectors which are perturbative from the orientifold
viewpoint \cite{GJ1}.
Then the gravitational anomaly cancellation condition implies that
after Higgsing the number of hypermultiplets in the open string sector is the
same in all three models (the closed string spectra of these models are
identical), hence they are on the same moduli. This implies that in the ${\bf
Z}_4$ and ${\bf Z}_6$ cases we also have $n_5=8$ and $n_I=16$. In the
orientifold language the corresponding Chan-Paton matrices are given by
\cite{GJ}:
\begin{eqnarray}\label{Z_4}
&&{\bf Z}_4:~~~\gamma_{g}={\mbox{diag}}(\beta {\bf I}_4,-\beta {\bf I}_4,
\beta^{-1}{\bf I}_4,-\beta^{-1}{\bf I}_4)~,\\
\label{Z_6}
&&{\bf Z}_6:~~~\gamma_{g}={\mbox{diag}}(i\alpha {\bf I}_2,-i\alpha {\bf I}_2,
i\alpha^{-1}{\bf I}_2,-i\alpha^{-1}{\bf I}_2, i{\bf I}_4,-i{\bf I}_4)~,
\end{eqnarray}
where $\beta\equiv\exp(\pi i/4)$.
{}Thus, the gauge bundles that are perturbative from the orientifold viewpoint
have $(n_5,n_I)=(0,24)$ and $(n_5,n_I)=(8,16)$. It is instructive to understand
these gauge bundles a bit better. In particular, here we would like to discuss
them in the heterotic and F-theory pictures. On the heterotic side there are
three topologically distinct gauge bundles with
$(n_5,n_I)=(0,24)$. These are characterized by the generalized second
Stiefel-Whitney class ${\widetilde w}_2$ (which is an element of
$H^2({\mbox{K3}},{\bf Z})$, or, more precisely, of
$H^2({\mbox{K3}},{\bf Z}_2)$ as ${\widetilde w}_2$ is defined modulo a shift
by twice a lattice vector of $H^2({\mbox{K3}},{\bf Z})$) \cite{berkooz,paul}.
The three distinct gauge bundles correspond to: (1) ${\widetilde w}_2=0$; (2)
${\widetilde w}_2\cdot{\widetilde w}_2=0~({\mbox{mod}}~4)$; (3)
${\widetilde w}_2\cdot{\widetilde w}_2=2~({\mbox{mod}}~4)$. In the first case
we have vector structure, while (2) and (3) are the cases without vector
structure.
{}To understand the above three cases better, it is useful to think about
the corresponding gauge bundles in terms of the ${\bf Z}_2$ orbifold action
on the ${\mbox{Spin}}(32)/{\bf Z}_2$ degrees of freedom in the heterotic
context. This action can be viewed in terms of a ${\bf Z}_2$ valued shift $V$
of the ${\mbox{Spin}}(32)/{\bf Z}_2$ lattice.
Let us use the Cartan basis of $SO(32)$ to
write $V$. If $2V$ belongs to the $SO(32)$ lattice,
then we have vector
structure, and ${\widetilde w}_2=0$. Two inequivalent gauge bundles of this
type which are perturbative from the heterotic viewpoint are given by
$V=((1/2)^2~0^{14})$ and $V=((1/2)^6~0^{10})$ (here we have chosen the
surviving spinor of ${\mbox{Spin}}(32)/{\bf Z}_2$ to be given by
$((1/2)^{16})$ modulo $SO(32)$ shifts). In particular, they satisfy the modular
invariance constraints such as the level matching condition. (Thus,
the gauge bundle given by $V=((1/2)^2~0^{14})$ corresponds to the standard
embedding for the ${\bf Z}_2$ orbifold.) These gauge bundles are
non-perturbative from the orientifold viewpoint.
{}However, we can have $V$ such that
$2V$ does not belong to the $SO(32)$ lattice, albeit it does belong to the
${\mbox{Spin}}(32)/{\bf Z}_2$ lattice. Two inequivalent gauge bundles of this
type are given by $V=((1/4)^{16})$ and $V=((1/4)^{15}(3/4))$. Note that
for such gauge bundles ${\widetilde w}_2\not=0$, and ${\widetilde w}_2\cdot
{\widetilde w}_2=(2V)^2$, where by $(2V)^2$ we mean the length squared of the
16-vector $2V$. The gauge bundle given by $V=((1/4)^{16})$ is
non-perturbative from the heterotic viewpoint - it does not satisfy the level
matching condition. In fact, this is precisely the gauge bundle in the
Type IIB orientifold of $T^4/{\bf Z}_2$ \cite{PS,GP}. As we have already
mentioned, this gauge bundle corresponds to embedding 16 instantons accompanied
by 8 5-branes to cancel the anomalies. On the other hand, the gauge bundle
$V=((1/4)^{15}(3/4))$ is perturbative from the heterotic viewpoint - it
satisfies the level matching requirement, and corresponds to embedding all 24
instantons. It is, however, non-perturbative from the orientifold viewpoint.
{}Finally, let us discuss the F-theory duals of the above cases. In \cite{SS}
it was shown that in the ${\widetilde w}_2=0$ case the dual F-theory
compactification is on the Calabi-Yau three-fold given by the $T^2$ fibration
over the base ${\bf F}_4$. (Here ${\bf F}_n$ are Hirzebruch surfaces.) In the
case of the ${\bf Z}_3$ orbifold this can be seen in two ways. First, we can
Higgs the $U(8)\otimes SO(16)$ gauge group at the orbifold point (see Table I)
down to $SO(8)\otimes SO(8)$ with 10 adjoint hypermultiplets in the first
$SO(8)$ and no matter in the second $SO(8)$. This is precisely the spectrum
of the F-theory dual on the Calabi-Yau three-fold with the base ${\bf F}_4$
(at the Voisin-Borcea orbifold point - see section IV).
On the other hand, we can
explicitly construct the $SO(8)\otimes SO(8)$ point in the string language.
Let K3 be $(T^2\otimes T^2)/{\bf Z}_3$, where ${\bf Z}_3$ simultaneously
rotates the two $T^2$'s by $2\pi/3$. Then consider the following choice of the
gauge bundle. First, let us turn on a ${\bf Z}_2$ valued
Wilson line on the $a$-cycle of, say, the first $T^2$ such that it breaks
$SO(32)$ down to $SO(16)\otimes SO(16)$. Next, let us turn on the second
${\bf Z}_2$ valued
Wilson line on the $b$-cycle of the same $T^2$ such that it further breaks
$SO(16)\otimes SO(16)$ down to $SO(8)^4$. It is not difficult to see that
this theory has ${\bf Z}_3$ symmetry under which, say, the first three
$SO(8)$'s are cyclically permuted. In fact, since the $T^2$'s are hexagonal,
under the ${\bf Z}_3$ rotation the $a$-cycles maps to the $b$-cycle, and so
the first Wilson line maps to the second Wilson line. The choice of the gauge
bundle corresponding to the ${\bf Z}_3$ orbifold then precisely consists of
the twist permuting the first three $SO(8)$'s, while leaving the last $SO(8)$
untouched. The resulting model has $SO(8)_3\otimes SO(8)_1$ gauge group, where
the subscript indicates the current algebra level via which the corresponding
subgroup is realized.
The untwisted sector contains the gravity supermultiplet, one tensor
supermultiplet, 2 neutral hypermultiplets, and one hypermultiplet in the
adjoint of $SO(8)_3$. The twisted sector contains 18 neutral hypermultiplets
(corresponding to the blow-up modes of the orbifold), and 9 hypermultiplets
in the adjoint of $SO(8)_3$. This spectrum precisely matches that of the
F-theory dual corresponding to the base ${\bf F}_4$.
{}As to the cases with
${\widetilde w}_2\cdot{\widetilde w}_2=0~({\mbox{mod}}~4)$ and
${\widetilde w}_2\cdot{\widetilde w}_2=2~({\mbox{mod}}~4)$, the corresponding
dual F-theory compactifications are on Calabi-Yau three-folds given by $T^2$
fibrations over ${\bf F}_0$ respectively ${\bf F}_1$ \cite{SS}. Here we note
that these cases also correspond to Voisin-Borcea orbifolds, but at the latter
points in the respective moduli spaces the corresponding bases are singular
leading to additional matter \cite{MV} precisely such that the resulting
massless spectra match those of the heterotic/Type I duals. Also, as was
pointed out in \cite{KST}, in the
${\widetilde w}_2\cdot{\widetilde w}_2=0~({\mbox{mod}}~4)$ case one can also
view the dual F-theory compactification as on a singular Calabi-Yau three-fold
(with Hodge numbers $(h^{1,1},h^{2,1})=(3,51)$) given by the ``non-geometric''
${\bf Z}_2\otimes {\bf Z}_2$ orbifold of $T^2\otimes T^2\otimes T^2$ with
discrete torsion \cite{VW}. This F-theory compactification is in turn on the
same moduli as that on a smooth Calabi-Yau three-fold (with Hodge
numbers $(h^{1,1},h^{2,1})=(3,243)$) given by the $T^2$ fibration over the
base ${\bf P}^1\otimes {\bf P}^1$ (that is, the smooth ${\bf F}_0$ surface)
\cite{KST}.
{}To summarize, the gauge bundles which are perturbative from the orientifold
viewpoint are those with $(n_5,n_I)=(0,24)$, ${\widetilde w}_2=0$ (this is the
case for the gauge bundle in the ${\bf Z}_3$ case (\ref{Z_3})), and
$(n_5,n_I)=(8,16)$, ${\widetilde w}_2\cdot{\widetilde w}_2=0~({\mbox{mod}}~4)$
(this is the case for the gauge bundles in the ${\bf Z}_2$ (\ref{Z_2}),
${\bf Z}_4$ (\ref{Z_4}), and ${\bf Z}_6$ (\ref{Z_6}) cases).
\section{Non-perturbative K3 Orientifolds}
{}In the previous section we have discussed the gauge bundles in K3
compactifications of Type I which are perturbative from the orientifold
viewpoint. In this section we would like to discuss the explicit Type I
vacua arising upon compactifications on K3 with these gauge bundles. These
compactifications can be viewed as the $\Omega$ orientifolds of Type IIB
on (the corresponding orbifold limits of) K3. Thus, the ${\bf Z}_2$ case
with the gauge bundle given by (\ref{Z_2}) (the action of the orbifold group
on both D9- and D5-branes is the same) was discussed in \cite{PS,GP}. This
model is completely perturbative from the orientifold viewpoint. Thus, the
``$\Omega$-twisted'' sector of the orientifold can be viewed as the usual
(``untwisted'') 99 and 55 open string sectors. The ``$\Omega g$-twisted''
sector (with $g^2=1$ in this case) can be viewed as the usual (``untwisted'')
59 open string sector. These sectors are perturbative from the orientifold
point of view. The massless open spectrum of this model is given in Table I.
{}In the ${\bf Z}_3$ case we do have non-perturbative sectors, however. Thus,
the ``$\Omega$-twisted'' sector of the orientifold corresponds to the
untwisted 99 open string sector (there are no D5-branes in this case). The
``$\Omega g$- and $\Omega g^2$-twisted'' sectors (with $g^3=1$ in this case)
correspond to the ${\bf Z}_3$
``twisted'' 99 open string sector. This model has a
perturbative heterotic dual which enables one to obtain its massless spectrum
\cite{NP}, and the latter is given\footnote{For the sake of simplicity,
throughout this paper we omit the $U(1)$ charges. In the untwisted open string
sectors they are straightforward to determine from the standard Chan-Paton
charge assignments. In the twisted open string sectors, however, these
$U(1)$ charges are generically fractional (in the normalization where
the fundamental of $SU(N)\subset U(N)$ in the untwisted open string sectors has
the $U(1)$ charge $+1$). In certain cases one can determine these
charges using Type I-heterotic duality (as on the heterotic side these charges
can be computed perturbatively in the corresponding twisted sectors). In other
cases determining the twisted sector $U(1)$ charges can be more involved. Also,
some of the twisted closed string states can transform non-trivially under
gauge transformations of some of the $U(1)$'s. The corresponding $U(1)$ charges
(in the appropriate basis) can therefore be non-zero (and sometimes even
fractional).} in Table I.
{}In the ${\bf Z}_6$ case we do not have a perturbative heterotic dual as the
Type I model contains D5-branes. Nonetheless, it is still possible to
determine the spectrum of the model. Thus, the ${\bf Z}_3$ twisted 99
open string sector is the same as in the ${\bf Z}_3$ model with the projection
onto the ${\bf Z}_2$ invariant states (note that
${\bf Z}_6\approx{\bf Z}_3\otimes {\bf Z}_2$). In particular, the 9
fixed points in the original ${\bf Z}_3$ twisted sector are combined into 5
linear combinations invariant under ${\bf Z}_2$ plus 4 linear combinations
which pick up a minus sign under the ${\bf Z}_2$ action. Taking this into
account, it is straightforward to determine the ${\bf Z}_2$ invariant states
in the ${\bf Z}_3$ twisted 99 open string sector (see \cite{NP} for details).
The ${\bf Z}_3$ twisted 55 open string sector is a bit more subtle - here we
assume that all D5-branes are sitting on top of each other at the orientifold
5-plane located at the origin of K3 ($z_1=z_2=0$). This implies that the
D5-branes only feel the singularity in K3 located at the origin. That is,
the other 8 of the original 9 fixed points in the ${\bf Z}_3$ twisted sector
play no role in this discussion as the twisted 55 states arise due to the {\em
local} geometry near the origin. This results in the fact that the
corresponding multiplicity in the ${\bf Z}_3$ twisted 55 open string
sectors is 1. Finally, the ``$\Omega g$- and $\Omega g^5$-twisted'' sectors
(with $g^6=1$ in this case) correspond to the ${\bf Z}_3$ twisted 59 open
string sector. In this case local geometry once again determines the
multiplicity of states to be 1 (and also fixes their quantum numbers)
\cite{NP}. The massless spectrum of the ${\bf Z}_6$ model is summarized in
Table I. Note that the twisted 99 and 55 open string sectors no longer
exhibit the naive ``T-duality''. This is due to the fact that these sectors
do not arise via a straightforward orbifold reduction of the corresponding
($SO(32)$) gauge theory (with ${\cal N}=2$ supersymmetry in six dimensions).
More concretely, the T-duality transformation involves ${\bf Z}_2$ reflection
of the complex coordinates $z_1,z_2$. The ${\bf Z}_3$ singularities, however,
are {\em not} invariant under this ${\bf Z}_2$ action, hence the lack of
T-duality symmetry.
{}Finally, let us consider the ${\bf Z}_4$ case which was not discussed in
\cite{NP}. The main difficulty in this case is that (unlike in the ${\bf Z}_3$
and ${\bf Z}_6$ cases) this model has no perturbative heterotic dual, nor
do we have any sectors of the theory which can be understood by performing
a perturbative computation on the heterotic side. Fortunately, however, anomaly
cancellation requirements (which are quite constraining in six dimensions)
along with the relevant geometric picture allow us to determine the massless
spectrum in this case as well. Thus, the gravitational anomaly cancellation
condition reads:
\begin{equation}\label{anom}
n_H-n_V=273-29n_T~,
\end{equation}
where $n_H,n_V,n_T$ are the numbers of hypermultiplets, vector multiplets
and tensor multiplets, respectively. In this case we have $n_T=1$ tensor
multiplets and $n^c_H=20$ hypermultiplets from the closed string sector.
All the vector multiplets arise in the open string sector. In particular,
$n_V=256$ as the gauge group is $[U(8)\otimes U(8)]_{99}\otimes
[U(8)\otimes U(8)]_{55}$. The anomaly cancellation condition (\ref{anom}) then
implies that the total number of open string hypermultiplets must be: $n^o_H=
n_H-n^c_H=480$. The perturbative sectors, namely, the untwisted 99 plus 55 as
well as 59 sectors give rise to 368 hypermultiplets in this model (see
Table I). Thus,
$112=4\times 28$ hypermultiplets must come from the twisted open string
sectors. These are the ${\bf Z}_4$ twisted 99 and 55 open string sectors
corresponding to the ``$\Omega g$- and $\Omega g^3$-twisted'' sectors (with
$g^4=1$ in this case). Note that there is no ``twisted'' 59 sector in this
model.
{}From the anomaly cancellation we get a hint that the twisted 99 and 55 open
string sector hypermultiplets must transform in the antisymmetric
representation ${\bf 28}$ of the $SU(8)$ subgroups. In fact, since the two
$SU(8)$'s within, say, the 55 open string sector are on the equal footing,
we conclude that the only way we can cancel anomalies is by assuming that
in the twisted
55 sector we have one copy of twisted hypermultiplets in $({\bf 28},
{\bf 1})_{55}$ and $({\bf 1},{\bf 28})_{55}$, and similarly for
the twisted 99 sector.
In the 55 sector this is actually what one expects from the geometric picture:
just as in the ${\bf Z}_6$ case here we have all D5-branes on top of each other
at the orientifold 5-plane located at the origin of K3, so we expect only {\em
one} copy of non-perturbative twisted hypermultiplets arising due to the local
geometry at the vicinity of the origin (the latter being the relevant
singularity in K3). However, for the 99 sector the conclusion that we have only
one copy of twisted hypermultiplets might appear a bit puzzling - the D9-branes
wrap the entire K3, and naively one expects that they should ``feel'' all 4
${\bf Z}_4$ fixed points.
{}Let us try to understand this point a bit better. An important observation
here is that the naive T-duality {\em is} expected to be a symmetry of this
background - indeed, the ${\bf Z}_2$ reflection of the complex coordinates
$z_1,z_2$ involved in the T-duality transformation leaves the corresponding
${\bf Z}_4$ singularities invariant. Thus, the ${\bf Z}_4$
twisted 55 and 99 open string sectors should have T-duality symmetry. Then we
must explain how come in the 99 sector we do not get 4 copies of the
corresponding matter representations but only one. The point here is that we
expect massless non-perturbative states to arise only for true geometric
singularities. However, if there is a non-zero twisted $B$-field turned on
inside of the ${\bf P}^1$ associated with a given blow-up mode (that is, with
a given fixed point), then non-perturbative states do not arise \cite{aspin}.
Since we are dealing with the ${\bf Z}_4$ twisted sectors, the corresponding
twisted $B$-field can only take values $0,\pm 1/4,1/2$ (here we normalize the
$B$-field so that it is defined modulo a unit shift). The $B$-filed must be
zero at the origin. At the other three fixed points, however, it must be
non-zero. Note that it cannot be $\pm 1/4$ as this would not be invariant
under the orientifold action $\Omega$ (recall that the $B$-field is odd under
the action of $\Omega$). However, the $B$-field at these three fixed points can
take a half-integer value which (taking into account that it is only defined
modulo an integer shift) is invariant under the action of $\Omega$. Thus, to
obtain a consistent ${\bf Z}_4$ orbifold background (with all D5-branes
located at the origin of K3), we must turn on half-integer twisted $B$-field
at the other three fixed points. Then the massless spectrum of the ${\bf Z}_4$
model, which is summarized in Table I, is free of anomalies.
Here we should note that in the
${\bf Z}_3$ and ${\bf Z}_6$
models the twisted $B$-field (in the ${\bf Z}_3$ twisted sectors) must be zero
as non-zero values such as $\pm 1/3$ would not be invariant under the action of
$\Omega$. This is why in these models all singularities (that is, fixed points)
are relevant in the twisted 99 sector.
{}Here we would like to
point out that all three ${\bf Z}_2,{\bf Z}_4,{\bf Z}_6$
models are on the same moduli. As we have already mentioned, at generic points
with completely broken gauge symmetry these models correspond to F-theory
compactifications on the Calabi-Yau three-fold (with
the Hodge numbers $(h^{1,1},h^{2,1})=(3,243)$) given by the $T^2$ fibration
over the base ${\bf P}^1\otimes {\bf P}^1$.
\section{Type I on Voisin-Borcea Orbifolds}
{}In this section we consider four dimensional ${\cal N}=1$ supersymmetric
Type I compactifications on Calabi-Yau three-folds known as Voisin-Borcea
orbifolds \cite{Voisin,Borcea}.
To aid the presentation, in the next subsection we review various
basic facts about Voisin-Borcea orbifolds following \cite{Asp,MV,KST}. To
clarify some points discussed in the previous sections, we also briefly discuss
F-theory compactifications on these spaces. Having reviewed the relevant
background material, we then discuss Type I compactifications on these
Calabi-Yau three-folds as well as their dual F-theory compactifications on
the corresponding Calabi-Yau four-folds known as Borcea four-folds.
\subsection{Voisin-Borcea Orbifolds}
{}Let ${\cal W}_2$ be a K3 surface (which is not necessarily an
orbifold of $T^4$) which admits an involution $J$ such that it reverses the
sign of the holomorphic
two-form $dz_1\wedge dz_2$ on ${\cal W}_2$. Consider the following quotient:
\begin{equation}
{\cal Y}_3= (T^2\otimes {\cal W}_2)/Y~,
\end{equation}
where $Y=\{1,S\}\approx {\bf Z}_2$, and $S$ acts as $Sz_0=-z_0$ on $T^2$
($z_0$ being a complex coordinate on $T^2$), and as $J$ on ${\cal W}_2$. This
quotient is a Calabi-Yau three-fold with $SU(3)$ holonomy
which is elliptically
fibered over the base ${\cal B}_2={\cal W}_2/B$, where
$B=\{1,J\}\approx{\bf Z}_2$.
{}Nikulin gave a classification \cite{Nik}
of possible involutions of K3 surfaces in terms of three
invariants $(r,a,\delta)$
(for a physicist's discussion, see, {\em e.g.}, \cite{Asp,MV}).
The result of this classification is plotted in Fig.2 (which has been borrowed
from \cite{KST})
according to the values of $r$
and $a$. The open and closed circles correspond to the cases with $\delta=0$
and $\delta=1$, respectively. (The cases denoted by ``$\otimes$'' are outside
of Nikulin's classification, and we will discuss them shortly.) In the case
$(r,a,\delta)=(10,10,0)$ the base ${\cal B}_2$ is an Enriques surface, and the
corresponding ${\cal Y}_3$ has Hodge numbers $(h^{1,1},h^{2,1})=(11,11)$.
In all the other cases the Hodge numbers are given by:
\begin{eqnarray}\label{hodge1}
&&h^{1,1}=5+3r-2a~,\\
\label{hodge2}
&&h^{2,1}=65-3r-2a~.
\end{eqnarray}
{}For $(r,a,\delta)=(10,10,0)$ the ${\bf Z}_2$ twist $S$ is freely acting
(that is, it has no fixed points).
For $(r,a,\delta)=(10,8,0)$ the fixed point set
of $S$ consists of two curves of genus 1. The base ${\cal B}_2$ in this case
is ${\bf P}^2$ blown up at 9 points. In all the other cases the fixed point set
of $S$ consists of one curve of genus $g$ plus $k$ rational curves, where
\begin{eqnarray}
&&g={1\over 2}(22-r-a)~,\\
&&k={1\over 2}(r-a)~.
\end{eqnarray}
{}It is sometimes useful to separate the above Hodge numbers into the
contribution $(h^{1,1}_0,h^{2,1}_0)$ of the untwisted sector and the
contribution $(h^{1,1}_*,h^{2,1}_*)$ of the ${\bf Z}_2$ twisted (that is,
$S$-twisted) sector. These are given by
\begin{eqnarray}\label{hodge10}
&&h^{1,1}_0=r+1~,\\
\label{hodge20}
&&h^{2,1}_0=21-r~,
\end{eqnarray}
and
\begin{eqnarray}\label{hodge1*}
&&h^{1,1}_*=4(k+1)~,\\
\label{hodge2*}
&&h^{2,1}_*=4g~,
\end{eqnarray}
respectively.
{}Note that except for the cases with $a=22-r$,
$r=11,\dots,20$, the mirror pair
of ${\cal Y}_3$ is given by the Voisin-Borcea orbifold
${\widetilde {\cal Y}}_3$ with ${\widetilde r}=20-r$, ${\widetilde a}=a$.
Under the mirror transform we have: ${\widetilde g}=f$, ${\widetilde f}=g$,
where $f=k+1$.
{}In the cases $a=22-r$, $r=11,\dots,20$, the mirror would have to have
${\widetilde r}=20-r$ and ${\widetilde a}=a={\widetilde r}+2$, where
${\widetilde r}=0,\dots,9$. We have depicted these cases in Fig.2 using
the ``$\otimes$'' symbol. In
particular,
we have plotted cases with $a=r+2$, $r=0,\dots,10$.
The Hodge numbers for these
cases are still given by (\ref{hodge1}) and (\ref{hodge2})
(which follows from
their definition as mirror pairs of the cases with $a=22-r$,
$r=11,\dots,20$). (This is
true for $a=r+2$, $r=0,\dots,9$.
Extrapolation to $r=10$ is motivated by
the fact that
in this case we get $(h^{1,1},h^{2,1})=(11,11)$ which is the same as for
$(r,a,\delta)=(10,10,0)$.) In \cite{KST} it was argued that these Voisin-Borcea
orbifolds also exist, albeit they are {\em singular}.
In fact, some of them can be constructed explicitly (see
\cite{KST} for details).
{}As we already mentioned, here we would like to briefly review
F-theory compactifications on Voisin-Borcea orbifolds (which correspond to
${\cal N}=1$ supersymmetric vacua in six dimensions). Thus, consider F-theory
on ${\cal Y}_3$ with\footnote{In the case $(r,a,\delta)=(10,10,0)$
we have $T=9$ tensor multiplets, $H=12$ hypermultiplets, and no gauge
vector multiplets. In the case $(r,a,\delta)=(10,8,0)$
we have $T=9$ tensor multiplets, $H=12$ neutral hypermultiplets,
gauge vector multiplets corresponding to $SO(8)\otimes SO(8)$, one
hypermultiplet in the adjoint of the first $SO(8)$, and one hypermultiplet in
the adjoint of the second $SO(8)$.} $(r,a,\delta)\not=
(10,10,0)$ or $(10,8,0)$. This
gives rise to the following massless spectrum in six
dimensions. The number of tensor multiplets is $T=r-1$. The number of neutral
hypermultiplets is $H=22-r$. The gauge group is $SO(8)\otimes SO(8)^k$.
There are $g$ adjoint hypermultiplets of the first $SO(8)$. There are no
hypermultiplets charged under the other $k$ $SO(8)$'s. Under mirror symmetry
$g$ and $f=k+1$ are interchanged. Thus, the vector multiplets in the adjoint
of $SO(8)^k$ are traded for $g-1$ hypermultiplets in the adjoint of the first
$SO(8)$. That is, gauge symmetry turns
into global symmetry and {\em vice-versa}.
This can be pushed further to understand F-theory compactifications on
Calabi-Yau three-folds
with $a=r+2$, $r=1,\dots,10$, which give the following
spectra\footnote{Here we
must exclude the case with $r=0$, $a=2$ for the F-theory prediction
would be $T=-1$
tensor multiplets. This Calabi-Yau three-fold
does exist \cite{KST},
but it is singular and F-theory compactification on this space does
not appear to
have a local Lagrangian description. However,
an extremal transition \cite{MV}
between this Calabi-Yau
three-fold and another Voisin-Borcea
orbifold could lead to a phase transition into
a well defined vacuum.}. The number of tensor multiplets is $T=r-1$. There are
$H=22-r$ neutral hypermultiplets. In
addition there are $g=10-r$ hypermultiplets
transforming as adjoints under a global $SO(8)$ symmetry.
There are no gauge bosons,
however. It is not difficult to verify that this massless
spectrum is free of gravitational
anomalies in six dimensions. In fact, F-theory compactifications
on these singular spaces are equivalent to F-theory compactifications
on smooth Calabi-Yau three-folds according to the following relation
\cite{KST}:
\begin{eqnarray}
&&{\mbox{F-theory on
${\cal Y}_3$ with $(h^{1,1},h^{2,1})=(r+1,61-5r)$ is equivalent
to}}\nonumber\\
&&{\mbox{F-theory on
${\widehat{\cal Y}}_3$ with $({\hat h}^{1,1},{\hat h}^{2,1})
=(r+1,301-29r)$ ($r=1,\dots,9$)}}~.
\end{eqnarray}
Thus, for instance,
for $r=2$ we get $({\hat h}^{1,1},{\hat h}^{2,1})=(3,243)$,
which is the elliptic Calabi-Yau given by the $T^2$ fibration over
the base ${\bf P}^1\otimes {\bf P}^1$. We have already
encountered this Calabi-Yau three-fold in the previous sections. Note that
F-theory compactifications on Voisin-Borcea orbifolds with $(r,a,\delta)=
(2,2,0),(2,2,1)$ are on the same moduli as the background corresponding
to $({\hat h}^{1,1},{\hat h}^{2,1})=(3,243)$. The reason is that the
corresponding bases in the former two cases (namely, the Hirzebruch surfaces
${\bf F}_0$ and ${\bf F}_1$) are singular at the Voisin-Borcea points
\cite{MV}.
After the appropriate blow-ups additional matter arises precisely such that
the moduli count matches that for the $({\hat h}^{1,1},{\hat h}^{2,1})=
(3,243)$ compactification.
\subsection{Type I on Voisin-Borcea Orbifolds}
{}In this subsection we consider some general aspects of Type I
compactifications on Voisin-Borcea orbifolds reviewed in the previous
subsection. In discussing such a compactification, we must specify the gauge
bundle. We discussed perturbative (from the orientifold viewpoint)
gauge bundles for K3 compactifications in section II. Our goal here will be
to determine which gauge bundles are perturbative from the orientifold
viewpoint for the Voisin-Borcea compactifications. It turns out that one can
answer this question using the map between Type IIB orientifolds and F-theory.
In fact, as we will see in a moment, the requirement that the gauge bundle
be perturbative greatly
restricts the possible choices of the compactification space itself.
{}Thus, consider Type I compactification on the Voisin-Borcea orbifold
${\cal Y}_3$ labeled by $(r,a,\delta)\not=(10,10,0)$. We can view this
as the $\Omega$ orientifold of Type IIB on ${\cal Y}_3$. This theory contains
32 D9-branes as well as D5-branes. Thus, there are D5-branes filling
${\bf R}^{3,1}\otimes {\cal C}$, where ${\bf R}^{3,1}$ is the non-compact four
dimensional Minkowski space-time, whereas ${\cal C}$ is the set of points
(of real dimension 2) in ${\cal B}_2$ fixed under the action of the
${\bf Z}_2$ twist $S$ (see the previous subsection for notations). There may be
other 5-branes depending upon the choice of the space ${\cal Y}_3$. Thus,
if the K3 surface ${\cal W}_2$ (recall that the base of the elliptic fibration
is ${\cal B}_2={\cal W}_2/J$) can be written as the quotient ${\cal W}_2=
{\cal W}_2^\prime/{\bf Z}_2$ (here ${\cal W}^\prime_2$ can be either $T^4$ or
another K3 surface), where the generator $R$ of ${\bf Z}_2$ reflects both of
the complex coordinates $z_1,z_2$ on ${\cal W}^\prime_2$, then we have other
5-branes in the theory. First, we have D5-branes filling ${\bf R}^{3,1}\otimes
{\cal C}^\prime$, where ${\cal C}^\prime$ is the set of points (of real
dimension 2) in ${\cal B}_2$ fixed under the action of the twist $SR$. Also,
there are 5-branes filling ${\bf R}^{3,1}\otimes T^2$ (here $T^2$ is the fibre
$T^2$). More precisely, as we will see in a moment,
this is the case for perturbative gauge bundles (while
for other choices of the gauge bundle these 5-branes may be absent).
{}To understand which gauge bundles are perturbative from the orientifold
viewpoint in the case of Voisin-Borcea compactifications, it is convenient
to T-dualize the corresponding Type I background in the direction of
the fibre $T^2$. This way we arrive at the $\Omega {\widetilde J}(-1)^{F_L}$
orientifold of Type IIB on ${\cal Y}_3$ (which contains various
7-branes plus, in certain cases, 3-branes). Here ${\widetilde J}$
reflects the
complex coordinate $z_0$ on the fibre $T^2$ (while acting trivially on the base
${\cal B}_2$), and $F_L$ is the left-moving
(space-time) fermion number. This orientifold can now be mapped to the
F-theory compactification on an elliptically fibered Calabi-Yau four-fold
via the map of \cite{sen} which we review next.
{}Consider an $\Omega {\widetilde J} (-1)^{F_L}$ orientifold of Type IIB on
${\cal Y}_3$, where ${\widetilde J}$
reverses the sign of the holomorphic 3-form
$dz_1\wedge dz_2\wedge dz_3$ on ${\cal Y}_3$, and the
set of points fixed under ${\widetilde J}$ in ${\cal Y}_3$ has real
dimension 4. Then following \cite{sen}
we can map this orientifold to (a limit of) F-theory on a
Calabi-Yau four-fold ${\cal X}_4$ defined as
\begin{equation}\label{CY_4}
{\cal X}_4=({\widetilde T}^2\otimes {\cal Y}_3)/X~,
\end{equation}
where $X=\{1,{\widetilde S}\}\approx{\bf Z}_2$, and ${\widetilde S}$ acts as
${\widetilde S} {\widetilde z}_0=-{\widetilde z}_0$ on ${\widetilde T}^2$
(${\widetilde z}_0$ is a complex coordinate on ${\widetilde T}^2$),
and as ${\widetilde J}$
on ${\cal Y}_3$. To avoid confusion, here we use
tilde to distinguish between the
fibre ${\widetilde T}^2$ in the elliptic fibration over the base
${\cal B}_3={\cal Y}_3/B$, where $B\equiv\{1,
{\widetilde J}\}\approx{\bf Z}_2$, and the fibre $T^2$ in the elliptic
fibration over the base ${\cal B}_2={\cal W}_2/Y$.
{}Note that in the case where ${\cal Y}_3$ is a Voisin-Borcea orbifold, the
above four-fold ${\cal X}_4$ can be viewed as a quotient $({\widetilde
{\cal W}}_2 \otimes {\cal W}_2)/Y$. Here ${\widetilde {\cal W}}_2\equiv
({\widetilde T}^2\otimes T^2)/X$ (recall that
$X=\{1,{\widetilde S}\}\approx{\bf Z}_2$, and ${\widetilde S}$ acts as
${\widetilde S}{\widetilde z}_0=-{\widetilde z}_0$ and ${\widetilde S}
z_0=-z_0$ on the complex coordinates parametrizing ${\widetilde T}^2$ and
$T^2$, respectively). Thus, ${\widetilde{\cal W}}_2$ is (the ${\bf Z}_2$
orbifold limit of) a K3 surface. The action of $Y=\{1,S\}$ on the two
K3 surfaces ${\widetilde {\cal W}}_2$ and ${\cal W}_2$ is given by
the corresponding Nikulin's involutions labeled by the integers
$({\widetilde r},{\widetilde a},{\widetilde \delta})$ and $(r,a,\delta)$
respectively. (In fact, it is not difficult to show that in the present case
${\widetilde r}=18$ and ${\widetilde a}=4$.) These Calabi-Yau four-folds
(which have $SU(4)$ holonomy) generically are singular spaces. The Euler
number $\chi$ of such a four-fold is given by \cite{Borcea}:
\begin{equation}
{1\over 24} \chi= 12+{1\over 4}({\widetilde r} -10) (r-10)~.
\end{equation}
Taking into account that in the case at hand we have ${\widetilde r}=18$,
we obtain
\begin{equation}
{1\over 24}\chi=12+2(r-10)=2(r-4)~.
\end{equation}
Now consider F-theory compactification on such a four-fold. To cancel
space-time anomaly we need to introduce $\chi/24$ 3-branes, which implies that
this number must be non-negative or else supersymmetry appears to be broken
\cite{SVW}. It then follows that
\begin{equation}\label{state}
{\mbox{Type I on Voisin-Borcea orbifolds can be tadpole free only for
$r\geq 4$.}}
\end{equation}
This is a non-trivial constraint with interesting implications for Type I
model building. For instance, if we take Type I on a ${\bf Z}_2\otimes
{\bf Z}_2$ orbifold of $T^2\otimes T^2\otimes T^2$ with discrete torsion, the
Hodge numbers are $(h^{1,1},h^{2,1})=(3,51)$ (note that this is the mirror
Calabi-Yau of a usual ${\bf Z}_2\otimes {\bf Z}_2$ orbifold without discrete
torsion whose Hodge numbers are $(h^{1,1},h^{2,1})=(51,3)$), and we have
$r=2$ and $a=4$. Thus, Type I compactification on this space is always
anomalous (that is, there is no choice of the gauge bundle consistent with
tadpole cancellation).
{}Thus, let us consider Type I compactifications with $r\geq 4$. In the dual
F-theory picture the number of 3-branes required for anomaly cancellation
is $2(r-4)$. If this number is non-zero, we have a choice of where to place
these 3-branes: ({\em i}) we can keep them in the bulk; from the Type I
point of view these correspond to dynamical 5-branes (made of some number of
D5-branes); ({\em ii})
alternatively, we can ``dissolve'' them into the 7-branes;
from the Type I viewpoint this corresponds to embedding a certain
${\mbox{Spin}}(32)/{\bf Z}_2$ gauge bundle. The corresponding instantons are no
longer point-like (at generic points). Thus, we see that we need to specify
additional data, which is
the choice of the gauge bundle. What we are interested
in understanding here is which choices are perturbative from the orientifold
viewpoint.
{}To answer this question recall that in K3 compactifications the total
number of instantons must be 24. By these we actually mean the number
$n_I$ of instantons embedded in the gauge bundle plus the number $n_5$ of
5-branes transverse to K3. In the case of the Voisin-Borcea compactifications
this number 24 is reduced to 12 - the corresponding pairing takes place
due to the additional ${\bf Z}_2$ projection by the orbifold group $Y$. Thus,
$\chi/24$ must be at least 12. This implies that the statement (\ref{state})
can be made even stronger:
\begin{equation}\label{state1}
{\mbox{Type I on Voisin-Borcea orbifolds can be tadpole free only for
$r\geq 10$.}}
\end{equation}
In fact, since we have a ${\bf Z}_2$ twist generated by $S$, we must include
32 D5-branes (filling ${\bf R}^{3,1}\otimes {\cal C}$ - see above).
These correspond to 8 dynamical 5-branes: here two pairings take place - one
due to the $\Omega$ orientifold projection, and the other one due to the ${\bf
Z}_2$ orbifold projection. Note that here we are making an assumption that
the corresponding background is perturbative from the orientifold viewpoint -
it is {\em a priori} not obvious at all that in other cases we must
include D5-branes. In fact, here we can give an example where this is {\em not}
the case. Thus, consider ${\mbox{Spin}}(32)/{\bf Z}_2$
heterotic on $T^4/{\bf Z}_2$ with the standard
embedding of the gauge bundle (which we discussed in section II). This theory
is perturbative from the heterotic viewpoint, hence no 5-branes are present.
That is, no D5-branes would be present on the dual Type I side either. This
compactification, however, is {\em non-perturbative} from the orientifold
viewpoint.
{}Thus, we must have at least 20 3-branes (12 from K3, and 8 from the D5-branes
filling ${\bf R}^{3,1}\otimes {\cal C}$). This implies that, for the gauge
bundle to be perturbative, we must have $r\geq 14$. In fact we will now show
that there are only two choices of $r$ for which we can have perturbative
gauge bundle: $r=14$ and $r=18$. First, suppose $r>14$. Then we must have
additional 5-branes whose number is $2(r-14)$. However, the only other way
we can obtain 5-branes perturbatively in the orientifold language is to
have 8 dynamical 5-branes (arising from 32 D5-branes) filling
${\bf R}^{3,1}\otimes {\cal C}^\prime$ (these are perpendicular to those
filling ${\bf R}^{3,1}\otimes {\cal C}$). This then implies that we must have
$r=18$. On the other hand, if $r=14$ we must make sure that there are no
5-branes filling ${\bf R}^{3,1}\otimes {\cal C}^\prime$. This implies that
in this case the K3 surface ${\cal W}_2$ {\em cannot} be written as the
quotient ${\cal W}^\prime/{\bf Z}_2$. It then follows that there are {\em no}
5-branes transverse to K3 (that is, ${\cal W}_2$) either (that is, $n_5=0$ and
$n_I=24$ in the corresponding K3 compactification).
{}The above discussion shows that the gauge bundle can be perturbative if and
only if $r=14$ or $r=18$. Let us apply these constraints to the cases where
the K3 surface ${\cal W}_2$ is a toroidal orbifold, that is, ${\cal W}_2=
T^4/{\bf Z}_N$ with $N=2,3,4,6$. In the $N=2$ case (without discrete torsion
between $S$ and the generator $g$ of the K3 orbifold group) the Hodge numbers
of the corresponding Voisin-Borcea orbifold are $(h^{1,1},h^{2,1})=(51,3)$, and
$r=18$, $a=4$. We then arrive at a consistent Type I compactification on
the ${\bf Z}_2\otimes {\bf Z}_2$ orbifold discussed in \cite{BL}. In the $N=3$
case the Hodge numbers are $(h^{1,1},h^{2,1})=(35,11)$, and we arrive at the
non-perturbative orientifold corresponding to Type on ${\bf Z}_6^\prime$
orbifold recently constructed in \cite{NP}. For later convenience we give
the massless spectrum of this model in Table II.
{}Next, consider the $N=4$ case. Suppose there is no discrete torsion between
$S$ and $g^2$ (where $g$ is the generator of the K3 orbifold group). Then the
Hodge numbers are $(h^{1,1},h^{2,1})=(61,1)$, and, therefore, $r=20$ and $a=2$.
On the other hand, if we include discrete torsion between $S$ and $g^2$, we
obtain $(h^{1,1},h^{2,1})=(3,51)$ just as in the ${\bf Z}_2\otimes {\bf Z}_2$
case with discrete torsion. Thus, in Type on the ${\bf Z}_2\otimes {\bf Z}_4$
orbifold we cannot cancel tadpoles if we choose the perturbative gauge bundle.
(In fact, in the case with
discrete torsion it is impossible to cancel tadpoles for any choice of the
gauge bundle. In the case without discrete torsion tadpole
cancellation might be possible
at the expense of choosing a gauge bundle which is non-perturbative from the
orientifold viewpoint.) This explains why the naive perturbative tadpole
cancellation conditions were found in \cite{Zw} to have no solution.
Here the map to F-theory provides a simple geometric interpretation of this
fact.
{}The above conclusion about perturbative inconsistency of Type I on
${\bf Z}_2\otimes {\bf Z}_4$ can be used to understand other similar
models. Note that the sectors of the orientifold labeled by $\Omega$,
$\Omega S$, $\Omega g^k$, $k=1,2,3$, and $\Omega Sg^2$ are perturbatively well
defined (as the corresponding K3 orientifolds are well defined). This implies
that the ``troublesome'' sectors are the $\Omega Sg$ and $\Omega Sg^3$ sectors.
This is precisely the conclusion that was reached in \cite{KST}.
Note that $Sg$ is the generator of an orbifold group ${\bf Z}_4$, which, to
avoid confusion, we will refer to as ${\bf Z}_4^*$. In particular,
$T^6/{\bf Z}_4^*$ is a Calabi-Yau three-fold with $SU(3)$ holonomy. Thus,
we conclude that Type I on the ${\bf Z}_4^*$ orbifold is not consistent for
perturbative gauge bundles. In fact,
we can further deduce that perturbative gauge bundles would lead to
inconsistencies in the ${\bf Z}_4\otimes {\bf Z}_4$, ${\bf Z}_8$ and
${\bf Z}_{12}^\prime$ orbifold models as well since all of these orbifold
groups contain ${\bf Z}_4^*$ as a subgroup. This is in accord with the
discussion in \cite{KST}, where it was also argued why in the
${\bf Z}_8^\prime$ model choosing the perturbative gauge bundle leads to
similar inconsistencies as well. Thus, the discussion in \cite{KST} explains
the results of \cite{ibanez}
(where it was shown that the naive perturbative tadpole cancellation conditions
in the ${\bf Z}_4^*$, ${\bf Z}_8$, ${\bf Z}_8^\prime$ and ${\bf Z}_{12}^\prime$
cases have no solution) using the geometric picture via
the map to F-theory. Here the ${\bf Z}_2\otimes {\bf Z}_4$ example serves
as an illustration of the discussion in \cite{KST}.
{}Finally, let us discuss the case $N=6$ (without discrete torsion between
$S$ and $g^3$, where $g$ is the generator of the K3 orbifold group). The
Hodge numbers in this case are $(h^{1,1},h^{2,1})=(51,3)$ just as in the
${\bf Z}_2\otimes {\bf Z}_2$ case without discrete torsion. This implies that
$r=18$ and $a=4$. This compactification, therefore, should be consistent
for the appropriate perturbative gauge bundle. We will
explicitly construct this model (via the corresponding non-perturbative
orientifold) in the next subsection.
\subsection{The ${\bf Z}_2\otimes {\bf Z}_6$ Model}
{}In this subsection we are going to give the massless spectrum of the
non-perturbative orientifold based on the ${\bf Z}_2\otimes {\bf Z}_6$ orbifold
compactification of Type I theory. This background has ${\cal N}=1$
supersymmetry in four dimensions. Here we should emphasize that the choice of
the gauge bundle in the model we discuss in this subsection is
{\em perturbative} from the orientifold viewpoint. The non-perturbative sectors
arise due to singularities in the Calabi-Yau itself as discussed in
Introduction. That is, non-perturbative states in this theory arise along the
lines of our discussion in section III for K3 compactifications of Type I.
{}Thus, consider Type I on the Voisin-Borcea orbifold ${\cal Y}_3=
(T^2\otimes T^2\otimes T^2)/({\bf Z}_2\otimes {\bf Z}_6)$, where
the generators $R_3$ and $g$ of ${\bf Z}_2$ respectively ${\bf Z}_6$ have
the following action on the complex coordinates $z_1,z_2,z_3$ parametrizing the
three two-tori:
\begin{eqnarray}
&&R_iz_j=-(-1)^{\delta_{ij}}z_j~,~i,j=1,2,3~,\\
&&\theta z_1=z_1~,~~~\theta z_2=\alpha z_2~,~~~\theta z_3=\alpha^{-1} z_3~,
\end{eqnarray}
where $R_1\equiv g^3$, $R_2\equiv R_1 R_3$, $\theta\equiv g^2$, and
$\alpha\equiv\exp(2\pi i/3)$. The Calabi-Yau three-fold ${\cal Y}_3$
(which has $SU(3)$ holonomy) has the Hodge numbers $(h^{1,1},h^{2,1})=
(51,3)$, so that there are $54$ chiral supermultiplets in the closed string
sector.
{}In the open string sector we have 32 D9-branes, and also three sets of
D5-branes with 32 D5-branes in each set. Thus, the world-volumes of the
D$5_i$-branes are four non-compact space-time dimensions plus the two-torus
parametrized by the complex coordinate $z_i$. For the sake of definiteness,
in the following we will concentrate on the brane configuration where all
D$5_i$-branes are placed on top of each other at the orientifold $5_i$-plane
located at the origin in the transverse dimensions (that is, $z_j=0$,
$j\not=i$).
{}Up to equivalent representations the Chan-Paton matrices are given by:
\begin{eqnarray}
&&\gamma_{\theta,9}={\mbox{diag}}({\bf W}\otimes {\bf I}_2,{\bf I}_8)~,\\
&&\gamma_{R_i,9}=i\sigma_i\otimes {\bf I}_8~.
\end{eqnarray}
Here ${\bf W}\equiv{\mbox{diag}}(\alpha,\alpha,\alpha^{-1},\alpha^{-1})$,
and $\sigma_i$ are the usual $2\times 2$ Pauli matrices. (The action on
the D$5_i$-branes is similar.) The perturbative (from the orientifold
viewpoint) massless open string spectrum of this model can be obtained
using the standard orientifold techniques, and is
given\footnote{Here we note that the perturbative spectrum of this model
was discussed in \cite{Zw}.}
in Table III.
Thus, the $\Omega$-twisted sector corresponds to the untwisted 99 plus
$5_i5_i$ open string sectors, while $\Omega R_i$ twisted sector
corresponds to the untwisted $95_i$ plus $5_j5_k$ ($j\not=k\not=i$) open string
sectors.
{}The non-perturbative sectors in this orientifold are the following.
The $\Omega\theta$- and $\Omega\theta^2$-twisted sectors
correspond to the twisted 99 plus $5_i5_i$ open string sectors, while
the $\Omega R_i\theta$- and $\Omega R_i\theta^2$-twisted sectors
correspond to the twisted $95_i$ plus $5_j5_k$ ($j\not=k\not=i$)
open string sectors. The twisted $95_1$ and $95_{2,3}$ sectors
are straightforward to work out
starting from the twisted 95 sector in the six dimensional
${\bf Z}_6$ model in Table I respectively the twisted 95 sector in the
four dimensional ${\bf Z}_6^\prime$ model in Table II and projecting onto the
corresponding ${\bf Z}_2$ invariant states. The twisted $5_25_3$ sector is
the same as the twisted $95_1$ sector (with the obvious substitutions of the
gauge quantum numbers). The twisted $5_1 5_{2,3}$ sectors are the same as
the twisted $9 5_{3,2}$ sectors except that the multiplicity of the states in
the latter is 1, while it is 3 in the former. This is due to the fact that
the the corresponding 5-branes only feel the local geometry in the vicinity
of the corresponding fixed points at which they are placed. (This is in
complete parallel with the corresponding discussion for the six dimensional
cases in \cite{NP} and section III.) As to the twisted $99$, $5_2 5_2$ and
$5_3 5_3$ sectors, they can be worked out by starting from the twisted
99 and 55 sectors in the ${\bf Z}_6^\prime$ model in Table II and projecting
onto the corresponding ${\bf Z}_2$ invariant states. Similarly, the twisted
$5_1 5_1$ sector (as well as the twisted 99 sector) can be worked out by
starting from the twisted 55 sector in the ${\bf Z}_6$ model in Table I and
projecting onto the corresponding ${\bf Z}_2$ invariant states.
The complete massless spectrum of the ${\bf Z}_2\otimes {\bf Z}_6$ model
including both perturbative and non-perturbative (from the orientifold
viewpoint) states is given in Table III. Note that all non-Abelian gauge
anomalies cancel in this model.
\section{Type I on Generalized Voisin-Borcea Orbifolds}
{}In the previous section we considered Type I compactifications on
Voisin-Borcea orbifolds of the form $(T^2\otimes{\mbox{K3}})/{\bf Z}_2$. Here
we are going to discuss Type I compactifications on generalized Voisin-Borcea
orbifolds of the form
\begin{equation}
{\cal Y}_3=(T^2\otimes {\cal W}_2)/{\bf Z}_N~,
\end{equation}
where ${\cal W}_2$ is a K3 surface, and the generator $\eta$ of ${\bf Z}_N$
acts as $\eta z_0=\omega z_0$ on the fibre $T^2$ (parametrized by the complex
coordinate $z_0$), and as $\eta \Omega_2=\omega^{-1}\Omega_2$ on ${\cal W}_2$
(parametrized by the complex coordinates $z_1,z_2$). Here $\omega\equiv
\exp(2\pi i/N)$, and $\Omega_2\equiv dz_1\wedge dz_2$ is the holomorphic
two-form on ${\cal W}_2$. Note that the holomorphic three-form $\Omega_3\equiv
dz_0\wedge dz_1\wedge dz_2$ on ${\cal Y}_3$ is invariant under the action
of $\eta$, which implies that ${\cal Y}_3$ is a Calabi-Yau three-fold with
$SU(3)$ holonomy (which is actually elliptically fibered over the base
${\cal W}_2/{\bf Z}_N$)
provided that the action of the orbifold group ${\bf Z}_N$ is a symmetry
of $T^2\otimes {\cal W}_2$. The latter requirement implies that $N$ is
restricted to the values $N=2,3,4,6$ so that the action of ${\bf Z}_N$ on $T^2$
is crystallographic. Note that in the $N=2$ case we have Voisin-Borcea
orbifolds discussed in the previous section. In this section we will be
interested in the cases $N=3,4,6$.
{}Let us start with the $N=4$ case. Suppose ${\cal W}_2$ is an orbifold K3.
Then the three-fold ${\cal Y}_3$ can be either the ${\bf Z}_4\otimes {\bf Z}_2$
or ${\bf Z}_4\otimes {\bf Z}_4$ orbifold of $T^2\otimes T^4$. We have already
discussed these cases in the previous section where we saw that for
perturbative gauge bundles tadpoles cannot be canceled in these models.
{}Next, let us consider the $N=6$ case with ${\cal W}_2$ an orbifold K3. Then
${\cal Y}_3$ can be the ${\bf Z}_6 \otimes {\bf Z}_2$, ${\bf Z}_6
\otimes {\bf Z}_3$, ${\bf Z}_6\otimes {\bf Z}_6$ or ${\bf Z}_6^*\otimes
{\bf Z}_2$ orbifold of
$T^2\otimes T^4$. The first case was explicitly constructed in the previous
section. We will discuss the last case, which was explicitly constructed
in \cite{223}, in a moment.
The other two cases are straightforward to work out: the perturbative
part of the spectrum is obtained using the standard orientifold techniques
\cite{Zw}, whereas the non-perturbative twisted open string sector states
can be easily worked out by starting with
the corresponding twisted states in the
${\bf Z}_6$ model in Table I, the ${\bf Z}_6^\prime$ model in Table II, and
the ${\bf Z}_2\otimes {\bf Z}_6$ model in Table III, and performing the
appropriate (${\bf Z}_2$ and/or ${\bf Z}_3$) orbifold projections.
Here the following remark is in order. Note that in the
${\bf Z}_6\otimes {\bf Z}_3$ and ${\bf Z}_6\otimes {\bf Z}_6$ models
{\em a priori} there are two types of twisted open string sectors. First, we
have open string sectors corresponding to the orientifold group elements
labeled by $\Omega\rho$, where $\rho$ is an orbifold group element such that
for some choice of $k\in 2{\bf Z}$ the set of points (in $T^2\otimes
{\cal W}_2$) fixed under $\rho^k$ is of real dimension 2. These are precisely
the twisted open string sectors which can be deduced via the appropriate
orbifold projections of the ``parent'' models mentioned above. This is due to
the fact that such twisted open string sectors are either projections of those
in the corresponding $T^2\otimes{\mbox{K3}}$ compactifications, or of
twisted 59
open string sectors which can be read off by appropriately projecting that in
the ${\bf Z}_6^\prime$ model in Table II. The second type of twisted open
string sectors correspond to the orientifold group elements labeled by $\Omega
\rho$ such that for some choice of $k\in 2{\bf Z}$ the set of points fixed
under $\rho^k$ is zero dimensional. Such twisted open string sectors were
argued in \cite{KST} to give rise to states that become heavy and decouple
upon appropriately blowing up the corresponding singularities in the
compactification space \cite{ZK,KS,KST}. We will discuss this point in more
detail shortly. Here, however, an important observation is that the second
type of twisted open string sectors do not give rise to massless states once we
consider {\em blown-up} orbifolds.
{}For the sake of brevity here we will not give the spectra of the
${\bf Z}_6\otimes {\bf Z}_3$ and ${\bf Z}_6\otimes {\bf Z}_6$
models. Let us mention,
however, that the gauge groups in these two cases are
$[U(2)^6\otimes U(4)]_{99}\otimes [U(2)^6\otimes U(4)]_{55}$ respectively
$[U(2)^3\otimes Sp(4)]_{99}\otimes \bigotimes_{i=1}^3
[U(2)^3\otimes Sp(4)]_{5_i5_i}$. Thus, the ${\bf Z}_6\otimes {\bf Z}_6$
model is non-chiral if we ignore the $U(1)$ charges. The
${\bf Z}_6\otimes {\bf Z}_3$ model is chiral as in the perturbative
open string sectors there are states transforming
under ${\bf 4}$ and ${\overline{\bf 4}}$ of the $SU(4)$ subgroups. However,
the twisted open string sector states are not charged under the $SU(4)$
subgroups, so that these sectors are automatically free of non-Abelian gauge
anomalies. These two models, therefore, are not as illuminating as, say, the
${\bf Z}_2\otimes {\bf Z}_6$ model in Table III.
{}Let us point out that the ${\bf Z}_6^*\otimes {\bf Z}_2$ model was explicitly
constructed in \cite{223}. The discussion of twisted open string sectors in
this model is completely analogous to that in the ${\bf Z}_6^*(\approx
{\bf Z}_3^*\otimes {\bf Z}_2)$ case. Let us therefore consider the generalized
Voisin-Borcea orbifolds with $N=3$ (and the ${\bf Z}_6^*$ case is a particular
example of this type).
{}For $N=3$ we have the following possibilities: ${\bf Z}_3^*\otimes {\bf Z}_2
\approx{\bf Z}_6^*$, ${\bf Z}_3^*\otimes {\bf Z}_4\approx{\bf Z}_{12}$,
${\bf Z}_3\otimes {\bf Z}_3$, and ${\bf Z}_3\otimes {\bf Z}_6$. The last case
we have already considered. The ${\bf Z}_3\otimes {\bf Z}_3$ case was
explicitly
constructed in \cite{KS}. Just as in the ${\bf Z}_6\otimes {\bf Z}_3$ and
${\bf Z}_6\otimes {\bf Z}_3$ cases, here we also have twisted open string
sectors of two types. As was shown in \cite{KS}, in this model twisted states
of {\em both} types decouple upon appropriately blowing up the orbifold
singularities, so that the remaining massless spectrum coincides with that
obtained in the perturbative orientifold approach. This turns out {\em not}
to be the case in the ${\bf Z}_{12}$ model which we will discuss in detail in
a moment. However, some of the sectors in the ${\bf Z}_{12}$ model are similar
to those in the ${\bf Z}_6^*$ model, which was explicitly constructed
in \cite{KS}, so let us discuss the latter in a bit more detail.
{}Thus, consider Type I on the generalized Voisin-Borcea orbifold
$(T^2\otimes T^2\otimes T^2)/{\bf Z}_6^*$, where the action of the
generator $g$ of ${\bf Z}_6^*$ on the complex coordinates $z_1,z_2,z_3$
parametrizing the three two-tori is given by:
\begin{equation}
gz_1=\alpha z_1~,~~~gz_{2,3}=-\alpha z_{2,3}~,
\end{equation}
where $\alpha\equiv\exp(2\pi i/3)$. In this case we have 32 D9-branes and 32
D5-branes, the latter wrapping the first $T^2$. The non-perturbative open
string sectors correspond to the orientifold group elements labeled by
$\Omega g$ plus $\Omega g^5$ and $\Omega g^2$ plus $\Omega g^4$. Note that
the singularities in the ${\bf Z}_6^*$ twisted sectors are a subset of
singularities in the ${\bf Z}_3^*$ twisted sectors. This implies that
if the blow-ups in the ${\bf Z}_3^*$ model lead to the decoupling of
non-perturbative open string sector states, the same should hold in the
${\bf Z}_6^*$ model as well. The fact that non-perturbative twisted open string
sector states indeed decouple in the blown-up ${\bf Z}_3^*$ model was shown
in \cite{ZK} using Type-I heterotic duality. Thus, in the corresponding ${\bf
Z}_6^*$ model we expect all non-perturbative states to decouple as well.
\subsection{The ${\bf Z}_{12}$ Model: A Non-perturbative Anomaly}
{}We are now ready to discuss the ${\bf Z}_{12}$ model. The generator $g$
of ${\bf
Z}_{12}$ has the following action on the three complex coordinates in the
compact directions:
\begin{equation}
gz_1=\alpha z_1~,~~~gz_2=i\alpha z_2~,~~~gz_3=-i\alpha z_3~.
\end{equation}
Note that ${\bf Z}_{12}\supset {\bf Z}_6^*$. The corresponding Calabi-Yau
four-fold has the Hodge numbers $(h^{1,1},h^{2,1})=(29,5)$, so that the closed
string spectrum contains 34 chiral supermultiplets. Here we note that the Hodge
numbers in the ${\bf Z}_6^*$ case are also $(h^{1,1},h^{2,1})=(29,5)$. That is,
in both the ${\bf Z}_6^*$ and ${\bf Z}_{12}$ models the compactification space
is the same, but the gauge bundles are different.
{}In the open string sector we have 32
D9-branes as well as 32 D5-branes (in the following we will focus on the
brane configuration where all D5-branes are sitting on top of each other
at the orientifold 5-plane located at $z_2=z_3=0$). The perturbative open
string sector of this model can be obtained using the standard orientifold
techniques (once the action of the orbifold group on the Chan-Paton factors
is specified - see below).
The non-perturbative sectors corresponding to the orientifold group
elements labeled by $\Omega g$ plus $\Omega g^{11}$, $\Omega g^2$ plus $\Omega
g^{10}$, $\Omega g^4$ plus $\Omega g^8$, and
$\Omega g^5$ plus $\Omega g^7$ do not give rise to massless states
(after the appropriate blow-ups) for the same reasons as in the ${\bf Z}_6^*$
case. However, the ${\bf Z}_4$ twisted open string sectors corresponding to the
orientifold group elements labeled by $\Omega g^3$ plus $\Omega g^9$ do give
rise to non-perturbative massless states. These states can be obtained by
starting from the ${\bf Z}_4$ model in Table I and projecting onto the
${\bf Z}_3^*$ invariant states. There is, however, a subtlety in this
projection, so let us discuss the action of the orbifold group on the
Chan-Paton factors in a bit more detail.
{}The point here is that to have a perturbative (from the orientifold
viewpoint) gauge bundle, we must make sure that all the twisted tadpole
cancellation conditions (derived in the perturbative orientifold approach)
cancel. In particular, the twisted tadpole cancellation condition for the
${\bf Z}_{12}$, ${\bf Z}_6^*$, ${\bf Z}_4$ and ${\bf Z}_2$ twisted Chan-Paton
matrices (for both D9- and D5-branes) read \cite{KS,ibanez}:
\begin{equation}
{\mbox{Tr}}(\gamma_{g^k})=0~,~~~k=1,2,3,5,6,7,9,10,11~.
\end{equation}
On the other hand, the ${\bf Z}_3^*$ twisted Chan-Paton matrices are fixed
by the corresponding tadpole cancellation conditions as in the ${\bf Z}_3^*$
model of \cite{Sagnotti} (see the next subsection for details).
The only way to satisfy all of these twisted
tadpole cancellation conditions is to consider $32\times 32$ Chan-Paton
matrices which (up to equivalent representations) are given by
\begin{eqnarray}
&&\gamma_{g^4}={\mbox{diag}}(\alpha{\bf I}_{12},\alpha^{-1}{\bf I}_{12},
{\bf I}_8)~,\\
&&\gamma_{g^3}={\mbox{diag}}({\bf U}\otimes {\bf I}_3,{\bf U}\otimes
{\bf I}_3,{\bf U}\otimes{\bf I}_2)~,
\end{eqnarray}
where ${\bf U}\equiv{\mbox{diag}}(\beta,-\beta,\beta^{-1},-\beta^{-1})$,
and $\beta\equiv\exp(\pi i/4)$. Note that the same would not be possible if
we considered $16\times 16$ matrices. That is, the tadpole cancellation
conditions require that both 16 D-branes and their orientifold images
participate in canceling the tadpoles at the same time,
while separate cancellations
of the corresponding tadpoles for the 16 D-branes and their orientifold images
cannot be achieved. This, in particular, implies that (since the 16 D-branes
and their orientifold images are mapped to each other by the orientifold
projection) the Chan-Paton matrix $\gamma_\Omega$ interchanges the twisted
Chan-Paton matrices with their conjugates: $\gamma_\Omega:\gamma_{g^k}
\rightarrow \gamma_{g^{-k}}$. Such an action is perfectly well defined
in the six dimensional $\Omega J^\prime$ orientifolds of \cite{GJ} as in those
cases the orientifold projection $\Omega J^\prime$ maps the $g$-twisted sector
to its conjugate $g^{-1}$-twisted sector.
However, as we have already pointed out in Introduction,
in four dimensions the orientifold projection is $\Omega$ which maps the
$g$-twisted sector to itself (here we assume that the appropriate blow-ups
have been performed). It is then {\em a priori} far from being obvious
whether it is consistent to have $\gamma_\Omega:\gamma_{g^k}
\rightarrow \gamma_{g^{-k}}$ while $\Omega: g^k\rightarrow g^k$.
In fact, we are going to argue that such an action leads to inconsistencies.
Here we should point out that such an inconsistency cannot be seen
perturbatively in the orientifold picture if we just focus on the gauge
(that is, the open string) sector of the theory. Indeed, the perturbative
open string sector of this model by itself is perfectly sensible (see below).
However, if one couples open and closed strings, then one expects an
inconsistency to show up - thus, the twisted closed string sector states
(which couple to the corresponding D-branes) feel the orientifold projection
in a way which is not compatible with the action of the orientifold projection
on the Chan-Paton factors. That is, the orientifold group actions on the closed
and open string degrees of freedom are not compatible. It is, however,
difficult to see this open-closed coupling inconsistency directly as the
$\Omega$ projection in the closed string sector makes sense only after the
appropriate blow-ups have been performed, so we are no longer at the orbifold
point in the corresponding
Calabi-Yau moduli space, which makes world-sheet computations
rather involved (and practically useless, at least for our purposes here).
{}There are, however, other ways to see this inconsistency. First, we can try
to understand it using Type I-heterotic duality. Here we should emphasize that
quantifying the following statements in this context is difficult as the
heterotic dual of the ${\bf Z}_{12}$ model would be non-perturbative - we have
5-branes in this case. Nonetheless, the following somewhat hand-waving argument
might still be useful. Thus, on the heterotic side the embedding of the gauge
bundle (that is, of the orbifold action on the gauge degrees of freedom)
can be understood in terms of the appropriate ${\bf Z}_{12}$ valued
${\mbox{Spin(32)}}/{\bf Z}_2$ lattice shifts. (These shifts are ${\bf Z}_{12}$
valued w.r.t. the ${\mbox{Spin(32)}}/{\bf Z}_2$ lattice, but are ${\bf Z}_{24}$
valued w.r.t. the $SO(32)$ lattice.) The $\Omega$ projection on the heterotic
side can be thought of as pairing 32 real Majorana fermions (in the real
fermion representation of the ${\mbox{Spin(32)}}/{\bf Z}_2$ degrees of freedom)
into 16 complex world-sheet fermions. The ${\bf Z}_{12}$ valued
${\mbox{Spin(32)}}/{\bf Z}_2$ lattice shifts in this language translate into
$U(1)^{16}$ phase rotations of the complex fermions. For these
rotations to be consistent, we must make sure that they can be written in the
complex $16\times 16$ basis. If this cannot be achieved, that is, if
we can only do this in the $32\times 32$ bases, then the model is
inconsistent\footnote{On the Type I side this would amount to an inconsistency
arising in the world-sheet theory of a probe D-string placed in this
background. I would like to thank E. Gimon for a discussion in a related but
somewhat different context.}.
In particular, the scattering amplitudes would be inconsistent \cite{KLST}.
In the following subsection we will relate this argument to the analogous
statement in the dual F-theory picture, where the above inconsistency can be
seen in a geometric way. However, for now let us take the above gauge bundle
and discuss the corresponding massless spectrum arising in the ${\bf Z}_{12}$
model. As we will see in a moment, the {\em non-perturbative} spectrum turns
out to be anomalous.
{}Once the twisted Chan-Paton matrices are fixed as above, the perturbative
(from the orientifold viewpoint) states in the ${\bf Z}_{12}$ model
are straightforward to work out.
The resulting perturbative
massless open string sector is given\footnote{Here we note that the
perturbative open string spectrum of the ${\bf Z}_{12}$ model was discussed
in \cite{ibanez}. The untwisted open string spectrum given in Table IV,
however, differs from that in \cite{ibanez}. In particular, note that the
spectrum should be invariant under permuting the two $U(2)$
subgroups (accompanied by an appropriate interchange/conjugation
of the corresponding
$U(3)$ subgroups) in, say, the 99 open string sector, which follows from the
action of the orbifold group on the Chan-Paton factors.
This is the case for the spectrum in Table IV, but {\em not} for the spectrum
given in \cite{ibanez}.} in Table IV. Note that the non-Abelian
gauge anomalies cancel in the untwisted open string
sector.
{}Next, let us discuss the non-perturbative open string sectors, that is, the
${\bf Z}_4$ twisted open string sectors. These are straightforward to work out
starting from the ${\bf Z}_4$ twisted open string sectors in the ${\bf Z}_4$
model in Table I and projecting onto the ${\bf Z}_3^*$ invariant states. Note
that the resulting twisted 99 and 55 open string sectors are apparently T-dual
for the same reason
as in the ${\bf Z}_4$ model in Table I - the ${\bf Z}_4$ fixed points are
invariant under the ${\bf Z}_2$ reflections involved in the T-duality
transformation (recall that the other twisted sectors such as
${\bf Z}_3^*$, ${\bf Z}_6^*$
and ${\bf Z}_{12}$ do not give rise to non-perturbative states once
the appropriate blow-ups are performed). In fact, it is important here that
the ${\bf Z}_4$ sector fixed point located at the origin contains no twisted
$B$-field, whereas the other three fixed points have half-integer valued
$B$-field stuck inside of the corresponding ${\bf P}^1$'s (see section III
for details). In particular, this configuration possesses the required
${\bf Z}_3$ symmetry.
{}Next, we list the resulting non-perturbative states arising in the
${\bf Z}_4$ twisted open string sectors (these states are ${\cal N}=1$
chiral supermultiplets - see Table IV for notations):
\begin{eqnarray}
&&({\overline {\bf 3}},{\bf 1},{\bf 1},{\bf 1},{\bf 1},
{\bf 1})_{99(T)/55(T)}~,\nonumber\\
&&({\bf 1},{\bf 1},{\overline {\bf 3}},{\bf 1},{\bf 2},
{\bf 1})_{99(T)/55(T)} ~,\nonumber\\
&&({\bf 1},{\bf 1},{\overline {\bf 3}},{\bf 1},{\bf 1},
{\bf 1})_{99(T)/55(T)}~,\nonumber\\
&&({\overline {\bf 3}},{\bf 1},{\bf 1},{\bf 1},{\overline {\bf 2}},
{\bf 1})_{99(T)/55(T)}~,\nonumber\\
&&({\bf 1},{\overline {\bf 3}},{\bf 1},{\bf 1},{\bf 1},
{\bf 1})_{99(T)/55(T)}~,\nonumber\\
&&({\bf 1},{\bf 1},{\bf 1},{\overline {\bf 3}},{\bf 1},
{\bf 2})_{99(T)/55(T)}~,\nonumber\\
&&({\bf 1},{\bf 1},{\bf 1},{\overline {\bf 3}},{\bf 1},
{\bf 1})_{99(T)/55(T)}~,\nonumber\\
&&({\bf 1},{\overline {\bf 3}},{\bf 1},{\bf 1},{\bf 1},
{\overline {\bf 2}})_{99(T)/55(T)}~,\nonumber
\end{eqnarray}
where the first four of these states come from the hypermultiplet
transforming in $({\bf 28},{\bf 1})$, and
the last four come from the hypermultiplet transforming in
$({\bf 1},{\bf 28})$ upon ${\bf Z}_3^*$ projecting the
$[U(8)\otimes U(8)]_{99/55}$ quantum numbers in the ${\bf Z}_4$ model in
Table I. Note that the above twisted spectrum suffers from $SU(3)^4$
non-Abelian gauge anomalies. That is, non-perturbatively we see an
inconsistency in this model. In the next subsection we will give additional
arguments clarifying the origins of this non-perturbative anomaly using the
map between Type IIB orientifolds and F-theory.
\subsection{Map to F-theory}
{}In the previous section we gave the map between Type I on Voisin-Borcea
orbifolds and F-theory. In this subsection we would like to do the same in the
case of generalized Voisin-Borcea orbifolds with $N=3,4,6$. There is a
subtlety in this map (as was pointed out in \cite{KST}) in the $N=3,6$ cases,
and here we will give make this map precise.
{}Naively, we can start from ${\cal Y}_3=(T^2\otimes {\cal W}_2)/{\bf Z}_N$,
T-dualize along the fibre $T^2$, and map the resulting model to F-theory
via \cite{sen} to arrive at the F-theory compactification on the following
Calabi-Yau four-fold:
\begin{equation}\label{four-fold}
{\cal X}_4=({\widetilde T}^2\otimes T^2\otimes {\cal W}_2)/
({\bf Z}_2\otimes {\bf Z}_N)~,
\end{equation}
where the action of the generator ${\widetilde S}$ of ${\bf Z}_2$ is given by
${\widetilde S}{\widetilde z}_0=-{\widetilde z}_0$, ${\widetilde S}z_0=-z_0$,
and it acts trivially on ${\cal W}_2$; the action of the generator $\eta$ of
${\bf Z}_N$ is given by $\eta {\widetilde z}_0={\widetilde z}_0$, $\eta z_0=
\omega z_0$, $\eta \Omega_2=\omega^{-1} \Omega_2$. Here ${\widetilde z}_0$ and
$z_0$ parametrize ${\widetilde T}^2$ respectively $T^2$, $\omega\equiv
\exp(2\pi i/N)$, and $\Omega_2$ is the holomorphic two-form on ${\cal W}_2$
(which is a K3 surface). Note, however, that the ${\bf Z}_N$
singularities in the ${\cal Y}_3$ fibre $T^2$ are invariant under
${\widetilde S}$ only in the $N=2,4$ cases. Thus, there is no subtlety in the
above map in these cases. In the $N=3,6$ cases, however, the ${\bf Z}_N$
singularities are not invariant under ${\widetilde S}$, so that we must be a
bit more careful. In fact, the issue in the $N=6$ case is the same as in the
$N=3$ case, so we can focus on the latter. Moreover, instead of considering
the generalized Voisin-Borcea orbifolds ${\cal Y}_3$, we will discuss this
point for a simpler example, namely, the ${\bf Z}_3^*$ orbifold. The
discussion for the other cases is completely analogous. Thus, let us start from
Type I on ${\cal Y}_3=(T^2\otimes T^2\otimes T^2)/{\bf Z}_3^*$, where the
generator $g$ of ${\bf Z}_3^*$ acts on the complex coordinates $z_1,z_2,z_3$
parametrizing the three two-tori as $gz_i=\alpha z_i$ ($i=1,2,3$, $\alpha\equiv
\exp(2\pi i/3)$). The Hodge numbers of this Calabi-Yau three-fold
are given by $(h^{1,1},h^{2,1})=(36,0)$.
{}In this Type I compactification, which was studied in detail
in \cite{Sagnotti,ZK,cvetic},
we have 32 D9-branes and no D5-branes. The twisted tadpole cancellation
conditions read ${\mbox{Tr}}(\gamma_g)=-2$ (in the $16\times 16$ basis). This
implies that (up to equivalent representations) we have
\begin{equation}
\gamma_g={\mbox{diag}}(\alpha{\bf I}_6,\alpha^{-1} {\bf I}_6,{\bf I}_4)~,
\end{equation}
where $\alpha\equiv\exp(2\pi i/3)$. The gauge group of this model is
$U(12)\otimes SO(8)$, and the (untwisted) 99 open string sector contains
the following chiral supermultiplets: $3\times ({\bf 66},{\bf 1})$ plus
$3\times ({\overline {\bf 12}},{\bf 8}_v)$.
{}Here we would like to ask what is the T-dual (where we T-dualize along both
directions in the first $T^2$) of this Type I background. That is, we would
like to understand the dual Type IIB orientifold with D7-branes. In fact, to
T-dualize we must first turn on Wilson lines which break the original $SO(32)$
gauge group (before ${\bf Z}_3^*$ orbifolding) to $SO(8)^4$. Indeed, the
corresponding T-dual of Type I on $T^2\otimes T^2\otimes T^2$ is the
$\Omega {\widetilde J} (-1)^{F_L}$ orientifold of Type IIB on $T^2\otimes
T^2\otimes T^2$, where ${\widetilde J} z_1=-z_1$, and ${\widetilde J} z_{2,3}=
z_{2,3}$. Note that $T^2/{\widetilde J}$ has four fixed points at which are
located the corresponding orientifold 7-planes. We must place 8 D7-branes at
each of these orientifold 7-planes to properly cancel tadpoles (this statement
is precise in the perturbative orientifold context). Thus, the gauge group
is $SO(8)^4$. In fact, let us see what are the locations of these fixed points.
In order for $T^2$'s to have ${\bf Z}_3^*$ symmetry (by which we are
ultimately going to mod out), they must be hexagonal. Let us focus on the first
$T^2$ whose volume we will denote by $v_1$. Then the metric on $T^2$ is given
by $g_{ab}=\sqrt{v_1/3} \,
e_a\cdot e_b$, $a,b=1,2$, where $e_a$ are the vectors
spanning the $SU(3)$ lattice. That is, $e_1^2=e_2^2=-2e_1\cdot e_2=2$. The
${\widetilde J}$ fixed points are located at $\xi_0=0$, $\xi_a=e_a/2$,
$a=1,2,3$, where $e_3\equiv -(e_1+e_2)$.
{}The T-dual Type I configuration of the above Type IIB orientifold is given
by the following. Start from Type I with 32 D9-branes on $T^2\otimes
T^2\otimes T^2$. Turn on two Wilson lines - one on the $a$-cycle and the other
one on the $b$-cycle of the first $T^2$ - such that they break the $SO(32)$
gauge symmetry down to $SO(8)^4$. This is precisely the configuration we are
looking for. Note that this setup is symmetric under ${\bf Z}_3$ valued
permutations of any three of the four $SO(8)$ subgroups accompanied by $2\pi/3$
rotations of the first $T^2$. Under these rotations the first Wilson line maps
to the second Wilson line (as the $a$-cycle maps to the $b$-cycle on $T^2$).
Therefore, we can mod out by the ${\bf Z}_3^*$ symmetry. The corresponding
gauge bundle is given by the following Chan-Paton matrix:
\begin{equation}
\gamma_g={\mbox{diag}}({\bf P}_3\otimes {\bf I}_4,{\bf W})~,
\end{equation}
where ${\bf P}_3$ is a $3\times 3$ matrix of cyclic permutations, and
${\bf W}\equiv{\mbox{diag}}(\alpha{\bf I}_2,\alpha^{-1}{\bf I}_2)$. Note that
we still have ${\mbox{Tr}}(\gamma_g)=-2$, so that the twisted tadpole
cancellation conditions are satisfied. The resulting gauge group is
$SO(8)\otimes U(4)$, and the matter consists of chiral supermultiplets in
$3\times ({\bf 28},{\bf 1})$ plus $3\times ({\bf 1},{\bf 6})$. The $SO(8)$
factor arises as the diagonal subgroup of the original $SO(8)^3$ subgroup
on which the permutation matrix ${\bf P}_3$ is acting. The $U(4)$ factor is
the corresponding subgroup of the fourth $SO(8)$.
{}The above $SO(8)\otimes U(4)$ model actually is on the same moduli as the
original $U(12)\otimes SO(8)$ model \cite{ZK}. The former point in the moduli
space is T-dual of the corresponding Type IIB orientifold with D7-branes, and
the latter can now be readily mapped to F-theory. The corresponding
compactification space is given by the four-fold ${\cal X}_4=
({\widetilde T}^2\otimes T^2\otimes T^2\otimes T^2)/({\bf Z}_2\otimes
{\bf Z}_3^*)$, which has $SU(4)$ holonomy, and is an elliptic fibration of
${\widetilde T}^2$ over the base $(T^2\otimes T^2\otimes T^2)/({\bf Z}_2
\otimes {\bf Z}_3^*)$.
{}We can generalize the above discussion to other examples of this type.
In particular, Type I on ${\cal Y}_3=(T^2\otimes {\cal W}_2)/{\bf Z}_N$
(where ${\cal W}_2$ is either a K3 surface or $T^4$) with the appropriately
turned on Wilson lines is dual to F-theory on the elliptic
four-fold (\ref{four-fold}). The Euler number of a four-fold is expressed in
terms of the corresponding Hodge numbers via:
\begin{equation}
\chi=4+2h^{1,1}-4h^{2,1}+2h^{3,1}+h^{2,2}~.
\end{equation}
However, for elliptic four-folds we have
\begin{equation}
h^{2,2}=44+4h^{1,1}-2h^{2,1}+4h^{3,1}~,
\end{equation}
so it suffices to give the Hodge numbers $(h^{1,1,},h^{2,1},h^{3,1})$ to
specify such a four-fold. In particular, we have
\begin{equation}
\chi/6=8+h^{1,1}-h^{2,1}+h^{3,1}~.
\end{equation}
For illustrative purposes, here we give
the Hodge numbers of the four-folds corresponding to Type I
on the ${\bf Z}_3^*$, ${\bf Z}_6^*$, ${\bf Z}_{12}^*$, and ${\bf Z}_6^*\otimes
{\bf Z}_2$ orbifolds (other cases are straightforward to work out):
\begin{eqnarray}
{\bf Z}_3^*:&~~~&(h^{1,1},h^{2,1},h^{3,1})=(32,21,9)~,~~~
\chi/24=7~,\nonumber\\
{\bf Z}_6^*,{\bf Z}_{12}:&~~~&(h^{1,1},h^{2,1},h^{3,1})=(32,6,14)~,~~~
\chi/24=12~,\nonumber\\
{\bf Z}_6^*\otimes {\bf Z}_2:&~~~&(h^{1,1},h^{2,1},h^{3,1})=(60,1,1)~,~~~
\chi/24=17~.\nonumber
\end{eqnarray}
Note that in the ${\bf Z}_6^*$ and ${\bf Z}_{12}$ cases we have the same
four-fold. This is not surprising as the corresponding three-folds
were also the same.
{}Finally, we would like to apply the above
map to F-theory to better understand the
non-perturbative inconsistency we have encountered in the ${\bf Z}_{12}$ case
with the particular gauge bundle discussed in the previous subsection. The
point here is that in the F-theory language the action of the orientifold
projection $\Omega$ is geometrized - it is mapped to the action of
${\widetilde J}$. Now, in the case of the four-fold compactifications of
F-theory we must specify the action of ${\widetilde J}$ not only on the
four complex coordinates, but also on the gauge bundle. More
precisely, in, say, the ${\bf Z}_{12}$ case we have to specify the action of
the orbifold group ${\bf Z}_{12}$ on various gauge degrees of freedom, that is,
we have to specify the gauge bundle. In the orientifold language this is
described by the Chan-Paton matrices, while in the heterotic language this is
done in terms of the ${\mbox{Spin(32)}}/{\bf Z}_2$ lattice shifts. In the
F-theory context we must also specify the orbifold action of the corresponding
gauge degrees of freedom. Thus, in the ${\bf Z}_3^*$ example discussed above
the ${\bf Z}_3^*$ orbifold acts geometrically on the three $SO(8)$'s
(corresponding to $D_4$ singularities in the fibre ${\widetilde T}^2$) by
permuting them - this is precisely what happens to the three ${\widetilde J}$
fixed points $\xi_a=e_a/2$, $a=1,2,3$, under the action of ${\bf Z}_3^*$.
However, the action of ${\bf Z}_3^*$ on the fourth $SO(8)$
(corresponding to the fixed point $\xi_0=0$) is no longer purely
geometric - it breaks $SO(8)$ down to $U(4)$, which implies that there is
non-trivial gauge bundle associated with the embedding of the ${\bf Z}_3^*$
orbifold action on the corresponding gauge degrees of freedom. The situation
here is similar to what happens in the heterotic compactifications - we must
embed some number of instantons which break the gauge group. Now, the action
of ${\widetilde J}$ on the ${\bf Z}_{12}$ twisted sectors {\em as well as}
the corresponding gauge bundles must be one and the same representation
of the orbifold group. In the gauge bundle we discussed in the previous
subsection the orbifold group (in the four-fold context) is ${\bf Z}_2\otimes
{\bf Z}_{12}$, whereas its embedding in terms of the gauge bundle would be
the non-Abelian dihedral group $D_{12}$. These two actions are clearly
incompatible, hence the non-perturbative inconsistency (which, in particular,
manifests itself via a non-perturbative anomaly) in the ${\bf Z}_{12}$ model.
\section{Extensions}
In the previous sections we have discussed various non-perturbative
orientifolds corresponding to Type I compactifications on K3 and Calabi-Yau
three-folds. Our discussion so far has been confined to the cases with trivial
NS-NS antisymmetric $B$-field. It would be interesting to understand
non-perturbative orientifolds with non-zero $B$-field, and one example of such
a compactification was recently discussed in \cite{NP}. We will not consider
these models in detail in this paper as we are planning to discuss such
compactifications elsewhere \cite{new}. However, here we would like to point
out an additional set of models arising in the {\em perturbative} K3
orientifolds (which complement those discussed in \cite{bij}) as their
generalizations to non-perturbative orientifolds might lead to interesting new
models in six and four dimensions.
{}The key point here is the following. Consider the $\Omega$ orientifold of
Type IIB on $T^4/{\bf Z}_2$ without the $B$-field. This is the model of
\cite{PS,GP}. Note that the action of the orbifold group on the Chan-Paton
factors is given by the Chan-Paton matrix
\begin{equation}
\gamma_g={\mbox{diag}}(i{\bf I}_8,-i{\bf I}_8)~,
\end{equation}
where $g$ is the generator of ${\bf Z}_2$. This corresponds to the gauge bundle
without vector structure (see section II for details). Here we can ask whether
we can take the Chan-Paton matrix to be given by
\begin{equation}
\gamma_g={\mbox{diag}}({\bf I}_8,-{\bf I}_8)~,
\end{equation}
which would also satisfy the twisted tadpole cancellation conditions. In this
case we would have the gauge bundle with vector structure. However, this choice
can be seen to be inconsistent. The point is that the gauge group in this
case is $[Sp(16)\otimes Sp(16)]_{99}\otimes [Sp(16)\otimes Sp(16)]_{55}$
(in our conventions $Sp(2N)$ has rank $N$), and the matter consists of
hypermultiplets in
\begin{eqnarray}
&& ({\bf 16},{\bf 16};{\bf 1},{\bf 1})_{99}~,~~~
({\bf 1},{\bf 1};{\bf 16},{\bf 16})_{55}~,\nonumber\\
&&{1\over 2}({\bf 16},{\bf 1};{\bf 16},{\bf 1})_{59}~,~~~
{1\over 2}({\bf 1},{\bf 16};{\bf 1},{\bf 16})_{59}~.\nonumber
\end{eqnarray}
Note that in the 59 sectors we have half-hypermultiplets in {\em real}
representations, which is inconsistent, albeit the gravitational anomaly
cancellation condition (\ref{anom}) is satisfied in this case (note that
the number of extra tensor multiplets is zero).
{}Note, however, that, as was pointed out in \cite{bij}, if we turn on the
$B$-field of rank $b$ ($b=2,4$ in the case of K3 compactifications), the 59
sector states come with multiplicity $2^{b/2}$ (while the rank of the gauge
group is reduced by $2^{b/2}$ \cite{Bij,bij}).
This implies that the 59 sector states no
longer need to come in half-hypermultiplets. Thus, consider the $b=2$
case. The Chan-Paton matrix is given by:
\begin{equation}\label{b=2}
\gamma_g={\mbox{diag}}({\bf I}_4,-{\bf I}_4)~.
\end{equation}
The gauge group now is $[Sp(8)\otimes Sp(8)]_{99}\otimes
[Sp(8)\otimes Sp(8)]_{55}$, and the matter consists of
hypermultiplets in
\begin{eqnarray}
&& ({\bf 8},{\bf 8};{\bf 1},{\bf 1})_{99}~,~~~
({\bf 1},{\bf 1};{\bf 8},{\bf 8})_{55}~,\nonumber\\
&&({\bf 8},{\bf 1};{\bf 8},{\bf 1})_{59}~,~~~
({\bf 1},{\bf 8};{\bf 1},{\bf 8})_{59}~.\nonumber
\end{eqnarray}
This spectrum is completely consistent, and, in particular, the gravitational
anomaly cancels (the number of extra tensor multiplets in this model is 4,
which follows from the results of \cite{bij}). Note that this model was
originally discussed in \cite{PS}.
{}In fact, we can generalize the above discussion to other perturbative
K3 orientifolds, that is, the $\Omega J^\prime$ orientifolds of Type IIB on
$T^4/{\bf Z}_N$, $N=2,3,4,6$. (Note that in the ${\bf Z}_2$ case the action of
$J^\prime$ is trivial.) Actually, in the ${\bf Z}_3$ case nothing changes
as we do not have a ${\bf Z}_2$ subgroup, but in other cases we do obtain new
models with non-zero $B$-field. Here we give the twisted Chan-Paton matrices
for these models, which we label $[N,b]$ ($g$ is the generator of
${\bf Z}_N$):
\begin{eqnarray}
&&[2,2]:~~~\gamma_g={\mbox{diag}}({\bf I}_4,-{\bf I}_4)~,\nonumber\\
&&[2,4]:~~~\gamma_g={\mbox{diag}}({\bf I}_2,-{\bf I}_2)~,\nonumber\\
&&[3,2]:~~~\gamma_g={\mbox{diag}}(\alpha{\bf I}_4,\alpha^{-1}
{\bf I}_4)~,\nonumber\\
&&[3,4]:~~~\gamma_g={\bf I}_4~,\nonumber\\
&&[4,2]:~~~\gamma_g={\mbox{diag}}(i{\bf I}_2,-i{\bf I}_2,{\bf I}_2,
-{\bf I}_2)~,\nonumber\\
&&[4,4]:~~~\gamma_g={\mbox{diag}}(i,-i,1,-1)~,\nonumber\\
&&[6,2]:~~~\gamma_g={\mbox{diag}}(\alpha {\bf I}_2,
-\alpha {\bf I}_2,\alpha^{-1} {\bf I}_2,-\alpha^{-1}{\bf I}_2)~,\nonumber\\
&&[6,4]:~~~\gamma_g={\mbox{diag}}({\bf I}_2,-{\bf I}_2)~,
\end{eqnarray}
where $\alpha\equiv\exp(2\pi i/3)$. These models, which were worked out in
\cite{unpublished}, are summarized in Table V. Note that in all of these
models the massless spectra satisfy the gravitational anomaly cancellation
condition (\ref{anom}). Here we note that the $[6,2]$ model in Table V is
the same as the corresponding $[6,2]$ model without vector structure
discussed in \cite{bij}. Also, the $[2,4]$ and $[6,4]$ models in Table V are
the same. Moreover, it is not difficult to show that each of the $[N,b]$ models
in Table V is on the same moduli as the corresponding $[N,b]$ model
without vector structure discussed in \cite{bij}.
\acknowledgments
{}This work was supported in part by the grant
NSF PHY-96-02074,
and the DOE 1994 OJI award. I would also like to thank Albert and
Ribena Yu for financial support.
\newpage
\begin{figure}[t]
\epsfxsize=16 cm
\epsfbox{figure.eps}
\bigskip
\caption{A schematic picture of the space of four dimensional
${\cal N}=1$ Type I and
heterotic vacua. The region ${\cal A}\cup{\cal B}$ corresponds to
perturbative Type I vacua.
The region ${\cal A}\cup{\cal C}$ corresponds to perturbative heterotic
vacua. The vacua
in the region ${\cal A}$ are perturbative from both the Type I and
heterotic viewpoints. The region ${\cal D}$ contains both non-perturbative
Type I and heterotic vacua.}
\end{figure}
\newpage
\begin{figure}
\setlength{\unitlength}{0.008in}%
$$\begin{picture}(445,266)(60,385)
\thinlines
\put(72,435){$\otimes$}
\put(92,455){$\otimes$}
\put(112,475){$\otimes$}
\put(132,495){$\otimes$}
\put(152,515){$\otimes$}
\put(172,535){$\otimes$}
\put(192,555){$\otimes$}
\put(212,575){$\otimes$}
\put(232,595){$\otimes$}
\put(252,615){$\otimes$}
\put(272,635){$\otimes$}
\put(100,420){\circle*{6}}
\put(140,420){\circle*{6}}
\put(260,420){\circle*{6}}
\put(300,420){\circle*{6}}
\put(420,420){\circle*{6}}
\put(460,420){\circle*{6}}
\put(120,440){\circle*{6}}
\put(160,440){\circle*{6}}
\put(240,440){\circle*{6}}
\put(280,440){\circle*{6}}
\put(320,440){\circle*{6}}
\put(400,440){\circle*{6}}
\put(440,440){\circle*{6}}
\put(140,460){\circle*{6}}
\put(180,460){\circle*{6}}
\put(220,460){\circle*{6}}
\put(260,460){\circle*{6}}
\put(300,460){\circle*{6}}
\put(340,460){\circle*{6}}
\put(380,460){\circle*{6}}
\put(420,460){\circle*{6}}
\put(160,480){\circle*{6}}
\put(200,480){\circle*{6}}
\put(240,480){\circle*{6}}
\put(280,480){\circle*{6}}
\put(360,480){\circle*{6}}
\put(400,480){\circle*{6}}
\put(180,500){\circle*{6}}
\put(220,500){\circle*{6}}
\put(260,500){\circle*{6}}
\put(300,500){\circle*{6}}
\put(340,500){\circle*{6}}
\put(380,500){\circle*{6}}
\put(200,520){\circle*{6}}
\put(240,520){\circle*{6}}
\put(280,520){\circle*{6}}
\put(320,520){\circle*{6}}
\put(360,520){\circle*{6}}
\put(220,540){\circle*{6}}
\put(260,540){\circle*{6}}
\put(300,540){\circle*{6}}
\put(340,540){\circle*{6}}
\put(240,560){\circle*{6}}
\put(280,560){\circle*{6}}
\put(320,560){\circle*{6}}
\put(260,580){\circle*{6}}
\put(300,580){\circle*{6}}
\put(280,600){\circle*{6}}
\put(320,480){\circle*{6}}
\put(120,400){\circle{10}}
\put(280,400){\circle{10}}
\put(440,400){\circle{10}}
\put(200,440){\circle{10}}
\put(200,480){\circle{10}}
\put(360,440){\circle{10}}
\put(360,480){\circle{10}}
\put(280,480){\circle{10}}
\put(280,520){\circle{10}}
\put(280,560){\circle{10}}
\put(280,600){\circle{10}}
\put(280,440){\circle{10}}
\put(120,440){\circle{10}}
\put(440,440){\circle{10}}
\put( 80,400){\line( 1, 0){420}}
\put( 80,400){\line( 0, 1){240}}
\put(300,620){\circle*{6}}
\put(320,600){\circle*{6}}
\put(340,580){\circle*{6}}
\put(380,540){\circle*{6}}
\put(400,520){\circle*{6}}
\put(420,500){\circle*{6}}
\put(440,480){\circle*{6}}
\put(440,480){\circle{10}}
\put(460,460){\circle*{6}}
\put(480,440){\circle*{6}}
\put(360,520){\circle{10}}
\put(360,560){\circle*{6}}
\put( 75,379){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{0}}}
\put(175,379){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{5}}}
\put(270,379){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{10}}}
\put(370,379){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{15}}}
\put(470,379){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{20}}}
\put( 65,495){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{5}}}
\put( 55,595){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{10}}}
\put(505,395){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{$r$}}}
\put( 75,645){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{$a$}}}
\end{picture}$$
\caption{Open circles and dots represent the original
Voisin--Borcea orbifolds.
The line of $\otimes$'s corresponds to the extension
discussed in section IV.}
\label{figVB}
\end{figure}
\begin{table}[t]
\begin{tabular}{|c|c|l|c|c|}
Model & Gauge Group & \phantom{Hy} Charged & Neutral
& Extra Tensor \\
& &Hypermultiplets & Hypermultiplets
&Multiplets \\
\hline
${\bf Z}_2$ & $U(16)_{99} \otimes U(16)_{55}$ &
$2 \times ({\bf 120};{\bf 1})_U$ & $20$
& $0$ \\
& & $2 \times({\bf 1};{\bf 120})_U$ & & \\
& & $({\bf 16};{\bf 16})_U$ & & \\
\hline
${\bf Z}_3$ & $[U(8) \otimes SO(16)]_{99}$ & $({\bf 28},{\bf 1})_U$ & $20$
& $0$ \\
& & $({\bf 8},{\bf 16})_U$ & & \\
& & $9\times ({\bf 28},{\bf 1})_T$ & & \\
\hline
${\bf Z}_4$ & $[U(8) \otimes U(8)]_{99}\otimes$
& $({\bf 28},{\bf 1};{\bf 1},{\bf 1})_U$ & $20$ & $0$ \\
& $[U(8) \otimes U(8)]_{55}$&
$({\bf 1},{\bf 28};{\bf 1},{\bf 1})_U$ & & \\
& & $({\bf 8},{\bf 8};{\bf 1},{\bf 1})_U$ & & \\
& & same as above with $99\leftrightarrow 55$ & & \\
& & $({\bf 8},{\bf 1};{\bf 8},{\bf 1})_U$ & & \\
& & $({\bf 1},{\bf 8};{\bf 1},{\bf 8})_U$ & & \\
& & $({\bf 28},{\bf 1};{\bf 1},{\bf 1})_T$ & & \\
& & $({\bf 1},{\bf 28};{\bf 1},{\bf 1})_T$ & & \\
& & $({\bf 1},{\bf 1};{\bf 28},{\bf 1})_T$ & & \\
& & $({\bf 1},{\bf 1};{\bf 1},{\bf 28})_T$ & & \\
\hline
${\bf Z}_6$ & $[U(4) \otimes U(4) \otimes U(8)]_{99}\otimes$
& $({\bf 6},{\bf 1},{\bf 1};{\bf 1},{\bf 1},{\bf 1})_U$ & $20$ & $0$ \\
& $[U(4) \otimes U(4) \otimes U(8)]_{55}$&
$({\bf 1},{\bf 6},{\bf 1};{\bf 1},{\bf 1},{\bf 1})_U$ & & \\
& & $({\bf 4},{\bf 1},{\bf 8};{\bf 1},{\bf 1},{\bf 1})_U$ & & \\
& & $({\bf 1},{\bf 4},{\bf 8};{\bf 1},{\bf 1},{\bf 1})_U$ & & \\
& & same as above with $99\leftrightarrow 55$ & &\\
& & $({\bf 4},{\bf 1},{\bf 1};{\bf 4},{\bf 1},{\bf 1})_U$ & & \\
& & $({\bf 1},{\bf 4},{\bf 1};{\bf 1},{\bf 4},{\bf 1})_U$ & & \\
& & $({\bf 1},{\bf 1},{\bf 8};{\bf 1},{\bf 1},{\bf 8})_U$ & & \\
& & $5\times ({\bf 6},{\bf 1},{\bf 1};{\bf 1},{\bf 1},{\bf 1})_T$ & & \\
& & $5\times ({\bf 1},{\bf 6},{\bf 1};{\bf 1},{\bf 1},{\bf 1})_T$ & & \\
& & $4\times ({\bf 4},{\bf 4},{\bf 1};{\bf 1},{\bf 1},{\bf 1})_T$ & & \\
& & $({\bf 1},{\bf 1},{\bf 1};{\bf 6},{\bf 1},{\bf 1})_T$ & & \\
& & $({\bf 1},{\bf 1},{\bf 1};{\bf 1},{\bf 6},{\bf 1})_T$ & & \\
& & $({\bf 4},{\bf 1},{\bf 1};{\bf 4},{\bf 1},{\bf 1})_T$ & & \\
& & $({\bf 1},{\bf 4},{\bf 1};{\bf 1},{\bf 4},{\bf 1})_T$ & & \\
\hline
\end{tabular}
\caption{The massless spectra of the six dimensional ${\cal N}=1$
supersymmetric Type IIB orientifolds
on $T^4/{\bf Z}_N$ for $N=2,3,4,6$.
The semi-colon in the column ``Charged Hypermultiplets'' separates $99$ and
$55$ representations. The subscript ``$U$'' indicates that the
corresponding (``untwisted'') state is perturbative from the orientifold
viewpoint. The subscript ``$T$'' indicates that the
corresponding (``twisted'') state is non-perturbative from the orientifold
viewpoint. The $U(1)$ charges are not shown, and by ``neutral''
hypermultiplets we mean that the corresponding states are not charged
under the non-Abelian subgroups.}
\end{table}
\begin{table}[t]
\begin{tabular}{|c|c|l|c|}
Model & Gauge Group & \phantom{Hy} Charged & Neutral
\\
& &Chiral Multiplets & Chiral Multiplets
\\
\hline
${\bf Z}_6^\prime$ & $[U(4) \otimes U(4) \otimes U(8)]_{99}\otimes$
& $({\bf 1},{\bf 1},{\bf 28};{\bf 1},{\bf 1},{\bf 1})_U$ & $46$ \\
& $[U(4) \otimes U(4) \otimes U(8)]_{55}$&
$({\bf 1},{\bf 1},{\overline {\bf 28}};
{\bf 1},{\bf 1},{\bf 1})_U$ & \\
& & $({\bf 4},{\bf 4},{\bf 1};{\bf 1},{\bf 1},{\bf 1})_U$ & \\
& & $({\overline {\bf 4}},{\overline {\bf 4}},{\bf 1};
{\bf 1},{\bf 1},{\bf 1})_U$ & \\
& & $({\bf 4},{\bf 1},{\bf 8};
{\bf 1},{\bf 1},{\bf 1})_U$ & \\
& & $({\bf 1},{\overline {\bf 4}},{\overline {\bf 8}};
{\bf 1},{\bf 1},{\bf 1})_U$ & \\
& & $({\overline {\bf 6}},{\bf 1},{\bf 1};
{\bf 1},{\bf 1},{\bf 1})_U$ & \\
& & $({\bf 1},{\bf 6},{\bf 1};
{\bf 1},{\bf 1},{\bf 1})_U$ & \\
& & $({\overline {\bf 4}},{\bf 1},{\bf 8};
{\bf 1},{\bf 1},{\bf 1})_U$ & \\
& & $({\bf 1},{\bf 4},{\overline {\bf 8}};
{\bf 1},{\bf 1},{\bf 1})_U$ & \\
& & $({\bf 4},{\overline {\bf 4}},{\bf 1};
{\bf 1},{\bf 1},{\bf 1})_U$ & \\
& & Same as above with $99\leftrightarrow 55$ & \\
& & $({\bf 4},{\bf 1},{\bf 1};{\bf 4},{\bf 1},{\bf 1})_U$ & \\
& & $({\bf 1},{\bf 4},{\bf 1};{\bf 1},{\bf 1},{\bf 8})_U$ & \\
& & $({\bf 1},{\bf 1},{\bf 8};{\bf 1},{\bf 4},{\bf 1})_U$ & \\
& & $({\overline {\bf 4}},{\bf 1},{\bf 1};{\bf 1},{\bf 1},
{\overline {\bf 8}})_U$ & \\
& & $({\bf 1},{\overline {\bf 4}},{\bf 1};{\bf 1},{\overline {\bf 4}},
{\bf 1})_U$ & \\
& & $({\bf 1},{\bf 1},{\overline {\bf 8}};{\overline {\bf 4}},{\bf 1},
{\bf 1})_U$ & \\
& & $6\times ({\bf 4},{\overline {\bf 4}},{\bf 1};
{\bf 1},{\bf 1},{\bf 1})_T$ & \\
& & $3\times ({\overline {\bf 4}},{\bf 4},{\bf 1};
{\bf 1},{\bf 1},{\bf 1})_T$ & \\
& & $6\times ({\overline {\bf 6}},{\bf 1},{\bf 1};
{\bf 1},{\bf 1},{\bf 1})_T$ & \\
& & $3\times ({\bf 6},{\bf 1},{\bf 1};
{\bf 1},{\bf 1},{\bf 1})_T$ & \\
& & $6\times ({\bf 1},{\bf 6},{\bf 1};
{\bf 1},{\bf 1},{\bf 1})_T$ & \\
& & $3\times ({\bf 1},{\overline {\bf 6}},{\bf 1};
{\bf 1},{\bf 1},{\bf 1})_T$ & \\
& & $3\times ({\bf 1},{\bf 1},{\bf 1};
{\bf 4},{\overline {\bf 4}},{\bf 1})_T$ & \\
& & $3\times ({\bf 1},{\bf 1},{\bf 1};
{\overline {\bf 6}},{\bf 1},{\bf 1})_T$ & \\
& & $3\times ({\bf 1},{\bf 1},{\bf 1};{\bf 1},{\bf 6},{\bf 1})_T$ & \\
& & $3\times ({\overline {\bf 4}},{\bf 1},{\bf 1};{\overline {\bf 4}},
{\bf 1},{\bf 1})_T$ & \\
& & $3\times ({\bf 1},{\bf 4},{\bf 1};{\bf 1},{\bf 4},{\bf 1})_T$ & \\
\hline
\end{tabular}
\caption{The massless spectrum of the four dimensional ${\cal N}=1$
supersymmetric Type IIB orientifold
on $T^6/{\bf Z}_6^\prime$.
The semi-colon in the column ``Charged Chiral Multiplets'' separates $99$ and
$55$ representations. The subscript ``$U$'' indicates that the
corresponding (``untwisted'') state is perturbative from the orientifold
viewpoint. The subscript ``$T$'' indicates that the
corresponding (``twisted'') state is non-perturbative from the orientifold
viewpoint. The $U(1)$ charges are not shown, and by ``neutral''
chiral multiplets we mean that the corresponding states are not charged
under the non-Abelian subgroups.}
\end{table}
\begin{table}[t]
\begin{tabular}{|c|c|l|c|}
Model & Gauge Group & \phantom{Hy} Charged & Neutral
\\
& &Chiral Multiplets & Chiral Multiplets
\\
\hline
${\bf Z}_2\otimes {\bf Z}_6$ & $[U(4) \otimes Sp(8)]_{99}\otimes$
& $({\bf 16},{\bf 1})_{99(U)/5_i5_i(U)}$ & $54$ \\
& $\bigotimes_{i=1}^3 [U(4) \otimes Sp(8)]_{5_i 5_i}$&
$({\bf 1},{\bf 28})_{99(U)/5_i5_i(U)}$ & \\
& & $({\bf 4}+{\overline {\bf 4}},{\bf 8})_{99(U)/5_i5_i(U)}$ & \\
& & $({\bf 6}+{\overline {\bf 6}},{\bf 1})_{99(U)/5_i5_i(U)}$ & \\
& & $({\bf 4},{\bf 1};{\overline {\bf 4}},{\bf 1})_{95_1(U)/5_2 5_3(U)}$ & \\
& & $({\overline {\bf 4}},{\bf 1};{\bf 4},{\bf 1})_{95_1(U)/5_2 5_3(U)}$ & \\
& & $({\bf 1},{\bf 8};{\bf 1},{\bf 8})_{95_1(U)/5_2 5_3(U)}$ & \\
& & $({\bf 4},{\bf 1};{\bf 4},{\bf 1})_{95_{2}(U)/5_1 5_{3}(U)}$ & \\
& & $({\overline {\bf 4}},{\bf 1};{\bf 1},{\bf 8})_{95_{2}(U)/5_1 5_{3}
(U)}$ & \\
& & $({\bf 1},{\bf 8};{\overline {\bf 4}},{\bf 1})_{95_{2}(U)/5_1 5_{3}
(U)}$ & \\
& & $({\overline {\bf 4}},{\bf 1};{\overline {\bf 4}},
{\bf 1})_{95_{3}(U)/5_1 5_{2}(U)}$ & \\
& & $({{\bf 4}},{\bf 1};{\bf 1},{\bf 8})_{95_{3}(U)/5_1 5_{2}
(U)}$ & \\
& & $({\bf 1},{\bf 8};{{\bf 4}},{\bf 1})_{95_{3}(U)/5_1 5_{2}
(U)}$ & \\
& & $2\times ({\bf 10}+{\overline {\bf 10}},{\bf 1})_{99(T)}$ & \\
& & $7\times ({\bf 6}+{\overline {\bf 6}},{\bf 1})_{99(T)}$ & \\
& & $({\bf 6}+{\overline {\bf 6}},{\bf 1})_{5_1 5_1(T)}$ & \\
& & $({\bf 10}/{\overline {\bf 10}},{\bf 1})_{5_2 5_2(T)/5_3 5_3(T)}$ & \\
& & $2\times ({\bf 6}/{\overline {\bf 6}},
{\bf 1})_{5_2 5_2(T)/5_3 5_3(T)}$ & \\
& & $3\times ({\overline {\bf 6}}/{\bf 6},
{\bf 1})_{5_2 5_2(T)/5_3 5_3(T)}$ & \\
& & $({\bf 4},{\bf 1};{\overline {\bf 4}},{\bf 1})_{95_1(T)/5_2 5_3(T)}$ & \\
& & $({\overline {\bf 4}},{\bf 1};{\bf 4},{\bf 1})_{95_1(T)/5_2 5_3(T)}$ & \\
& & $3\times ({\overline {\bf 4}},{\bf 1};{\overline {\bf 4}},
{\bf 1})_{95_{2}(T)}$
& \\
& & $3\times ({{\bf 4}},{\bf 1};{{\bf 4}},
{\bf 1})_{95_{3}(T)}$
& \\
& & $({{\bf 4}},{\bf 1};{{\bf 4}},
{\bf 1})_{5_1 5_{2}(T)}$
& \\
& & $({\overline {\bf 4}},{\bf 1};{\overline {\bf 4}},
{\bf 1})_{5_1 5_{3}(T)}$
& \\
\hline
\end{tabular}
\caption{The massless spectrum of the four dimensional ${\cal N}=1$
supersymmetric Type IIB orientifold
on $T^6/({\bf Z}_2\otimes {\bf Z}_6)$.
The semi-colon in the column ``Charged Chiral Multiplets'' separates $99$ and
the corresponding $5_i5_i$
representations. The subscript ``$U$'' indicates that the
corresponding (``untwisted'') state is perturbative from the orientifold
viewpoint. The subscript ``$T$'' indicates that the
corresponding (``twisted'') state is non-perturbative from the orientifold
viewpoint. The $U(1)$ charges are not shown, and by ``neutral''
chiral multiplets we mean that the corresponding states are not charged
under the non-Abelian subgroups. Note that ${\bf 16}$ and ${\bf 28}$ are
{\em reducible} representations of $SU(4)$ and $Sp(8)$, respectively
(in our conventions $Sp(2N)$ has rank $N$).}
\end{table}
\begin{table}[t]
\begin{tabular}{|c|c|l|c|}
Model & Gauge Group & \phantom{Hy} Charged & Neutral
\\
& &Chiral Multiplets & Chiral Multiplets
\\
\hline
${\bf Z}_{12}$ & $[U(3)^4 \otimes U(2)^2]_{99}\otimes$
& $({\bf 3},{\bf 1},{\bf 3},{\bf 1},{\bf 1},{\bf 1})_{99(U)/55(U)}$ & $34$ \\
& $[U(3)^4 \otimes U(2)^2]_{55}$&
$({\bf 1},{\bf 3},{\bf 1},{\bf 3},{\bf 1},{\bf 1})_{99(U)/55(U)}$ & \\
& & $({\overline {\bf 3}},{\bf 1},{\bf 1},{\bf 1},{\bf 2},
{\bf 1})_{99(U)/55(U)}$ & \\
& & $({\bf 1},{\overline {\bf 3}},{\bf 1},
{\bf 1},{\bf 1},{\bf 2})_{99(U)/55(U)}$ & \\
& & $({\bf 1},{\bf 1},{\overline {\bf 3}},
{\bf 1},{\overline {\bf 2}},{\bf 1})_{99(U)/55(U)}$ & \\
& & $({\bf 1},{\bf 1},{\bf 1},{\overline {\bf 3}},
{\bf 1},{\overline {\bf 2}})_{99(U)/55(U)}$ & \\
& & & \\
& & $({\overline {\bf 3}},{\bf 1},{\bf 1},{\bf 1},{\bf 1},
{\bf 1})_{99(U)/55(U)}$ & \\
& &
$({\bf 1},{\overline {\bf 3}},{\bf 1},{\bf 1},{\bf 1},
{\bf 1})_{99(U)/55(U)}$ & \\
& & $({\bf 1},{\bf 1},{\bf 3},{\bf 3},{\bf 1},
{\bf 1})_{99(U)/55(U)}$ & \\
& & $({\overline {\bf 3}},{\bf 1},{\bf 1},
{\bf 1},{\bf 1},{\overline {\bf 2}})_{99(U)/55(U)}$ & \\
& & $({\bf 1},{\overline {\bf 3}},{\bf 1},
{\bf 1},{\overline {\bf 2}},{\bf 1})_{99(U)/55(U)}$ & \\
& & $({\bf 1},{\bf 1},{\overline {\bf 3}},{\bf 1},
{\bf 2},{\bf 1})_{99(U)/55(U)}$ & \\
& & $({\bf 1},{\bf 1},{\bf 1},{\overline {\bf 3}},
{\bf 1},{\bf 2})_{99(U)/55(U)}$ & \\
& & &\\
& & $({\bf 1},{\bf 1},{\overline {\bf 3}},{\bf 1},{\bf 1},
{\bf 1})_{99(U)/55(U)}$ & \\
& &
$({\bf 1},{\bf 1},{\bf 1},{\overline {\bf 3}},{\bf 1},
{\bf 1})_{99(U)/55(U)}$ & \\
& & $({\bf 3},{\bf 3},{\bf 1},{\bf 1},{\bf 1},
{\bf 1})_{99(U)/55(U)}$ & \\
& & $({\overline {\bf 3}},{\bf 1},{\bf 1},
{\bf 1},{\overline {\bf 2}},{\bf 1})_{99(U)/55(U)}$ & \\
& & $({\bf 1},{\overline {\bf 3}},{\bf 1},
{\bf 1},{\bf 1},{\overline {\bf 2}})_{99(U)/55(U)}$ & \\
& & $({\bf 1},{\bf 1},{\overline {\bf 3}},{\bf 1},
{\bf 1},{\bf 2})_{99(U)/55(U)}$ & \\
& & $({\bf 1},{\bf 1},{\bf 1},{\overline {\bf 3}},
{\bf 2},{\bf 1})_{99(U)/55(U)}$ & \\
& & &\\
& & $({\bf 3},{\bf 1},{\bf 1},{\bf 1},
{\bf 1},{\bf 1};{\bf 1},{\bf 1},{\bf 3},{\bf 1},
{\bf 1},{\bf 1})_{59(U)}$ & \\
& & $({\bf 1},{\bf 3},{\bf 1},{\bf 1},
{\bf 1},{\bf 1};{\bf 1},{\bf 1},{\bf 1},{\bf 3},
{\bf 1},{\bf 1})_{59(U)}$ & \\
& & $({\bf 1},{\bf 1},{\bf 3},{\bf 1},
{\bf 1},{\bf 1};{\bf 3},{\bf 1},{\bf 1},{\bf 1},
{\bf 1},{\bf 1})_{59(U)}$ & \\
& & $({\bf 1},{\bf 1},{\bf 1},{\bf 3},
{\bf 1},{\bf 1};{\bf 1},{\bf 3},{\bf 1},{\bf 1},
{\bf 1},{\bf 1})_{59(U)}$ & \\
& & $({\overline {\bf 3}},{\bf 1},{\bf 1},{\bf 1},
{\bf 1},{\bf 1};{\bf 1},{\bf 1},{\bf 1},{\bf 1},
{\bf 2},{\bf 1})_{59(U)}$ & \\
& & $({\bf 1},{\overline {\bf 3}},{\bf 1},{\bf 1},
{\bf 1},{\bf 1};{\bf 1},{\bf 1},{\bf 1},{\bf 1},
{\bf 1},{\bf 2})_{59(U)}$ & \\
& & $({\bf 1},{\bf 1},{\overline {\bf 3}},{\bf 1},
{\bf 1},{\bf 1};{\bf 1},{\bf 1},{\bf 1},{\bf 1},
{\overline {\bf 2}},{\bf 1})_{59(U)}$ & \\
& & $({\bf 1},{\bf 1},{\bf 1},{\overline {\bf 3}},
{\bf 1},{\bf 1};{\bf 1},{\bf 1},{\bf 1},{\bf 1},
{\bf 1},{\overline {\bf 2}})_{59(U)}$ & \\
& & $({\bf 1},{\bf 1},{\bf 1},{\bf 1},
{\bf 2},{\bf 1};{\overline {\bf 3}},{\bf 1},{\bf 1},{\bf 1},
{\bf 1},{\bf 1})_{59(U)}$ & \\
& & $({\bf 1},{\bf 1},{\bf 1},{\bf 1},
{\bf 1},{\bf 2};{\bf 1},{\overline {\bf 3}},{\bf 1},{\bf 1},
{\bf 1},{\bf 1})_{59(U)}$ & \\
& & $({\bf 1},{\bf 1},{\bf 1},{\bf 1},
{\overline {\bf 2}},{\bf 1};{\bf 1},{\bf 1},{\overline {\bf 3}},{\bf 1},
{\bf 1},{\bf 1})_{59(U)}$ & \\
& & $({\bf 1},{\bf 1},{\bf 1},{\bf 1},
{\bf 1},{\overline {\bf 2}};{\bf 1},{\bf 1},{\bf 1},{\overline {\bf 3}},
{\bf 1},{\bf 1})_{59(U)}$ & \\
\hline
\end{tabular}
\caption{The {\em perturbative}
massless spectrum of the four dimensional ${\cal N}=1$
supersymmetric Type IIB orientifold
on $T^6/{\bf Z}_{12}$.
The semi-colon in the column ``Charged Chiral Multiplets'' separates $99$ and
$55$ representations. The subscript ``$U$'' indicates that the
corresponding (``untwisted'') state is perturbative from the orientifold
viewpoint. The $U(1)$ charges are not shown, and by ``neutral''
chiral multiplets we mean that the corresponding states are not charged
under the non-Abelian subgroups.}
\end{table}
\begin{table}[t]
\begin{tabular}{|c|c|c|l|c|c|}
Model & $b$ & Gauge Group & \phantom{Hy} Charged & Neutral
& Extra Tensor \\
& & &Hypermultiplets & Hypermultiplets
&Multiplets \\
\hline
${\bf Z}_2$ & 2 & $[Sp(8) \otimes Sp(8)]_{99}$ & $({\bf 8},{\bf 8};
{\bf 1},{\bf 1})$ & $16$ & $4$ \\
& &$[Sp(8) \otimes Sp(8)]_{55}$ & $({\bf 1},{\bf 1};{\bf 8},{\bf 8})$ & & \\
& & & $({\bf 8},{\bf 1};{\bf 8},{\bf 1})$ & & \\
& & & $({\bf 1},{\bf 8};{\bf 1},{\bf 8})$ & & \\
\hline
& 4 & $[Sp(4) \otimes Sp(4)]_{99}$ & $({\bf 4},{\bf 4};
{\bf 1},{\bf 1})$ & $14$ & $6$ \\
& &$[Sp(4) \otimes Sp(4)]_{55}$ & $({\bf 1},{\bf 1};{\bf 4},{\bf 4})$ & & \\
& & & $2\times ({\bf 4},{\bf 1};{\bf 4},{\bf 1})$ & & \\
& & & $2\times ({\bf 1},{\bf 4};{\bf 1},{\bf 4})$ & & \\
\hline
${\bf Z}_4$ & 2 & $[U(4) \otimes Sp(4) \otimes Sp(4)]_{99}\otimes$
&$({\bf 4},{\bf 4},{\bf 1};{\bf 1},{\bf 1},{\bf 1})$ & $14$ & $6$ \\
& & $[U(4) \otimes Sp(4) \otimes Sp(4)]_{55}$&
$({\bf 4},{\bf 1},{\bf 4};{\bf 1},{\bf 1},{\bf 1})$ & & \\
& & & $({\bf 1},{\bf 1},{\bf 1};{\bf 4},{\bf 4},{\bf 1})$ & & \\
& & & $({\bf 1},{\bf 1},{\bf 1};{\bf 4},{\bf 1},{\bf 4})$ & & \\
& & & $({\bf 4},{\bf 1},{\bf 1};{\bf 1},{\bf 4},{\bf 1})$ & & \\
& & & $({\bf 4},{\bf 1},{\bf 1};{\bf 1},{\bf 1},{\bf 4})$ & & \\
& & & $({\bf 1},{\bf 4},{\bf 1};{\bf 4},{\bf 1},{\bf 1})$ & & \\
& & & $({\bf 1},{\bf 1},{\bf 4};{\bf 4},{\bf 1},{\bf 1})$ & & \\
\hline
& 4 & $[U(2) \otimes Sp(2) \otimes Sp(2)]_{99}\otimes$
&$({\bf 2},{\bf 2},{\bf 1};{\bf 1},{\bf 1},{\bf 1})$ & $13$ & $7$ \\
& & $[U(2) \otimes Sp(2) \otimes Sp(2)]_{55}$&
$({\bf 2},{\bf 1},{\bf 2};{\bf 1},{\bf 1},{\bf 1})$ & & \\
& & & $({\bf 1},{\bf 1},{\bf 1};{\bf 2},{\bf 2},{\bf 1})$ & & \\
& & & $({\bf 1},{\bf 1},{\bf 1};{\bf 2},{\bf 1},{\bf 2})$ & & \\
& & & $2\times ({\bf 2},{\bf 1},{\bf 1};{\bf 1},{\bf 2},{\bf 1})$ & & \\
& & & $2\times({\bf 2},{\bf 1},{\bf 1};{\bf 1},{\bf 1},{\bf 2})$ & & \\
& & & $2\times({\bf 1},{\bf 2},{\bf 1};{\bf 2},{\bf 1},{\bf 1})$ & & \\
& & & $2\times ({\bf 1},{\bf 1},{\bf 2};{\bf 2},{\bf 1},{\bf 1})$ & & \\
\hline
${\bf Z}_6$ & 2 & $[U(4) \otimes U(4)]_{99}\otimes$
& $({\bf 6},{\bf 1};{\bf 1},{\bf 1})$ & $14$ & $6$ \\
& &$[U(4) \otimes U(4)]_{55}$& $({\bf 1},{\bf 6};{\bf 1},{\bf 1})$ & & \\
& & & $({\bf 4},{\bf 4};{\bf 1},{\bf 1})$
& & \\
& & & $({\bf 1},{\bf 1};{\bf 6},{\bf 1})$
& & \\
& & & $({\bf 1},{\bf 1};{\bf 1},{\bf 6})$
& & \\
& & & $({\bf 1},{\bf 1};{\bf 4},{\bf 4})$
& & \\
& & & $2\times ({\bf 4},{\bf 1};{\bf 4},{\bf 1})$
& & \\
& & & $2\times ({\bf 1},{\bf 4};{\bf 1},{\bf 4})$
& & \\
\hline
& 4 & $[Sp(4) \otimes Sp(4)]_{99}$ & $({\bf 4},{\bf 4};
{\bf 1},{\bf 1})$ & $14$ & $6$ \\
& &$[Sp(4) \otimes Sp(4)]_{55}$ & $({\bf 1},{\bf 1};{\bf 4},{\bf 4})$ & & \\
& & & $2\times ({\bf 4},{\bf 1};{\bf 4},{\bf 1})$ & & \\
& & & $2\times ({\bf 1},{\bf 4};{\bf 1},{\bf 4})$ & & \\
\hline
\end{tabular}
\caption{The massless spectra of the six dimensional Type IIB orientifolds
on $T^4/{\bf Z}_N$ for $N=2,4,6$, and various values of $b$
(the rank of the $B$-field) worked out in [35].
The semi-colon in the column ``Charged Hypermultiplets'' separates $99$ and
$55$ representations.}
\end{table}
|
\section{Introduction}
Domain walls and other solitons play a central role in many theories of
elementary particles. In 1983 Rubakov and Shaposhnikov (RS) suggested that
solitons might even constitute an alternative to compactification in
Kaluza-Klein and string models [1]. They gave as an example $\mathbf{\ \phi }
^{4}$-theory in five flat (non-compact) dimensions, choosing
\begin{equation}
\mathcal{L}=\tfrac{1}{2}\sum_{\mu =0}^{4}\left( \frac{\partial \mathbf{\phi }%
}{\partial \mathbf{x}^{\mathbf{\mu }}}\right) \left( \frac{\partial \mathbf{%
\phi }}{\partial \mathbf{x}_{\mathbf{\mu }}}\right) +\tfrac{1}{2}m^{2}%
\mathbf{\phi }^{2}-\tfrac{1}{4}\lambda \mathbf{\phi }^{4},
\end{equation}
with $\mathbf{\phi }$ a real, one-component field and $\mathbf{g}^{\mathbf{%
\mu \nu }}=$ diag$.\left( +,-,-,-;-\right) .$ If $m^{2}>0,$ then domain-wall
solutions can occur. RS noted that if a domain-wall appears and is assigned
to the fifth dimension, then an excitation of the wall can be interpreted as
a boson free to propagate in the four dimensions of ordinary space-time but
trapped on the wall in the extra dimension. What sets this manner of
confinement apart from ordinary compactification scenarios, of course, is
that a particle of sufficient energy might escape confinement and propagate
freely in all five dimensions.
RS also generated a massless (chiral) Dirac particle by adding to lagrangian
1 the terms
\begin{equation}
\mathcal{L}_{D}=i\overline{\mathbf{\psi }}\sum_{\mu =0}^{4}\mathbf{\gamma }^{%
\mathbf{\mu }}\partial _{\mathbf{\mu }}\mathbf{\psi }+g\overline{\mathbf{%
\psi }}\mathbf{\phi \psi ,}
\end{equation}
where $\ \mathbf{\gamma }^{\mathbf{\mu }}\mathbf{\gamma }^{\mathbf{\nu }}%
\mathbf{+\;\mathbf{\gamma }^{\mathbf{\nu }}\mathbf{\gamma }^{\mathbf{\mu }%
}=2g}^{\mathbf{\mu \nu }},\;\mathbf{\mu ,\nu }=0,1,\,.\,.\,4.$ \ As in the
case of the boson, the Dirac particle is free to propagate in $M^{4}$ but
skates on the wall in the fifth flat dimension.
Trapping a Dirac particle on an extra flat dimension also plays a role in
lattice gauge theories these days. To avoid Dirac-particle doubling, Kaplan
introduced a fifth flat dimension and generated a massless Dirac particle
with a lagrangian similar to that of Rubakov and Shaposhnikov [2]. When
Kaplan's lagrangian is expressed on a lattice, the second Weyl fermion
appears on an opposite wall in the fifth dimension with exponentially
vanishing overlap with the first Weyl fermion, thus getting around the no-go
theorems [3]. Jansen provides a review of work following Kaplan's original
paper [4].
A question that naturally comes to mind is whether there exists a soliton in
a larger space which can generate \textit{all three generations} of quarks
and leptons. RS remarked that if a domain wall could dynamically confine a
particle in one flat extra dimension, then perhaps a vortex could confine it
in two, a monopole in three, and an instanton in four flat extra dimensions.
Could the mass-spectrum of such a confined particle exhibit the quantum
numbers of the quarks and leptons?
For our part, we had carried out a study of the Klein-Gordon equation
extended to eight flat dimensions $M^{4}\times \widetilde{R}^{4}$ and found
that if we included a symmetrical harmonic-oscillator term in the four extra
dimensions, then we obtained confined $SU\left( 4\right) $ solutions
suggestive of the three generations of quarks and leptons [5]. However a
physical basis for the harmonic-oscillator term was lacking. Could a soliton
(instanton?) provide the confinement?
In the remainder of this paper, we propose a Dirac field equation in eight
flat dimensions which generates Dirac particles with quantum numbers
suggestive of quarks and leptons. We will indicate how these particles might
be coupled to gluons and the electroweak bosons\footnote{%
Dvali and Shifman have proposed a mechanism for localizing massless \emph{\
gauge} bosons on a domain wall [6]; they illustrate this with a domain wall
on the $(x,y)$ plane in $M^{4}$.} in a manner consistent with a left-right
symmetric extension of the Standard Model [7].
\section{Domain walls in five dimensions}
\smallskip To motivate the extension to eight dimensions, let us review the $%
\mathbf{\phi }^{4}$-model in five dimensions. Lagrangian 1 generates the
field equation\footnote{%
Symbols relating to $M^{4}$($\tilde{R}$ , $\,M^{4}\times \tilde{R}$ ) will
be printed normally (with a tilde, in \textbf{bold-face type}). $\tilde{R}$
refers to either $\tilde{R}^{1}$ or $\tilde{R}^{4}$.}
\begin{equation}
\sum_{\mathbf{\mu }=0}^{4}\partial ^{\mathbf{\mu }}\partial _{\mathbf{\mu }}%
\mathbf{\phi }-m^{2}\mathbf{\phi }+\lambda \mathbf{\phi }^{3}=0
\end{equation}
which admits the ``kink'' solution
\begin{equation}
\mathbf{\phi =}\left( m/\sqrt{\lambda }\right) \tanh \left[ m\left(
\widetilde{x}-\widetilde{x}_{0}\right) /\sqrt{2}\right] \equiv \widetilde{%
\phi }_{cl}.
\end{equation}
(Here we denote the fifth dimension $x^{4}\equiv \widetilde{x},$ and set $%
\widetilde{x}_{0}$ equal to zero.) $\widetilde{\phi }_{cl}$ minimizes the
action locally, and if $\mathbf{\phi }$ is expanded about it, then$\;\mathbf{%
\eta \equiv \phi -}\widetilde{\phi }_{cl}$ satisfies
\begin{equation}
\left( \sum_{\mathbf{\mu }=0}^{4}\partial ^{\mathbf{\mu }}\partial _{\mathbf{%
\mu }}-m^{2}+3\lambda \widetilde{\phi }_{cl}^{2}\right) \mathbf{\eta }=0;
\end{equation}
terms cubic and quartic in $\mathbf{\eta }$ are to be treated by standard
perturbation theory. If we set $\mathbf{\eta =}\eta \left( x^{\mu }\right)
\widetilde{\eta }\left( \widetilde{x}\right) ,$ then
\begin{equation}
\left( \square +M^{2}\right) \eta =0
\end{equation}
where $\square $ is the d'Alembertian in $M^{4}$, and
\begin{equation}
\left( -\partial ^{2}/\partial \widetilde{x}^{2}+3\lambda \widetilde{\phi }%
_{cl}^{2}-m^{2}\right) \widetilde{\eta }=M^{2}\widetilde{\eta }.
\end{equation}
$M$ is taken to be the scalar boson's mass. Eq. 7 is a one-dimensional
Schr\"{o}dinger-like equation that admits a massless solution (which merely
corresponds to a translation of the soliton), a solution $\widetilde{\eta }%
_{3/2}=\sinh \widetilde{z}/\cosh ^{2}\widetilde{z}$, ($\tilde{z}\equiv m%
\tilde{x}/\sqrt{2}$) with $M^{2}=\frac{3}{2}m^{2}$ which represents a boson
trapped in the well, and continuum solutions $\widetilde{\eta }_{c}$ for $%
M^{2}\geq 2m^{2}$ which represent bosons free to propagate in all five
dimensions [1, 8]. We have plotted $\widetilde{\eta }_{3/2}$ in Fig. 1 along
with the confining potential $3\lambda \widetilde{\phi }_{cl}^{2}-m^{2}%
\equiv \tilde{V}$.
\begin{figure}[h]
\epsfxsize=13cm \centerline{\epsfbox{fig1epsf.ps}}
\caption{Confining potential $\widetilde{V}$ as a function of the extra
dimension $\widetilde{x}$;$\;$\newline
$\widetilde{V}$ is generated by domain-wall solution $\widetilde{\phi }_{cl}$%
. Also shown is the wavefunction $\widetilde{\eta }_{3/2}$ of the boson
trapped in the well, and the wavefunction $\widetilde{\psi }_{0}$ of a
massless Dirac particle trapped on the wall, for $g=\protect\sqrt{\lambda }.$%
}
\end{figure}
If lagrangian 1 is augmented by lagrangian 2, then minimizing the action
yields the field equation
\begin{equation}
-i\sum_{\mathbf{\mu }=0}^{4}\mathbf{\gamma }^{\mathbf{\mu }}\partial _{%
\mathbf{\mu }}\mathbf{\psi }-g\mathbf{\phi \psi }=0.
\end{equation}
If $\mathbf{\phi }$ is approximated by the domain-wall solution $\widetilde{
\phi }_{cl}$, then
\begin{equation}
-i\sum_{\mathbf{\mu }=0}^{4}\mathbf{\gamma }^{\mathbf{\mu }}\partial _{%
\mathbf{\mu }}\mathbf{\psi }-g\widetilde{\phi }_{cl}\mathbf{\psi }=0.
\end{equation}
This equation admits a solution $\mathbf{\psi =}\psi _{0,L}\widetilde{\psi }
_{0}$, where $\psi _{0,L}$ is a left-helical, massless wavefunction in $%
M^{4} $ and
\begin{equation}
\widetilde{\psi }_{0}=\left( \cosh m\widetilde{x}/\sqrt{2}\right) ^{-g\sqrt{
2/\lambda }}.
\end{equation}
$\widetilde{\psi }_{0}$ is plotted in Fig 1 for the case $g=\sqrt{\lambda }.$
(There is no massless right-helical solution.) There are also unconfined
Dirac states of mass $\geq $ $gm/\sqrt{\lambda }.$
\section{Eight-dimensional Dirac equation}
Let us suppose that there exists a generalization of Eq. 9 in a
higher-dimensional flat space of sufficient dimensionality to accommodate
all of the quarks and leptons in the Standard Model. If so, then how many
extra dimensions are required, and what kind of ``potential'' is needed? The
difficulty is that $\widetilde{N}$-dimensional extensions naturally lead to $%
SO(\widetilde{N})$-symmetry groups, whereas quarks require $SU(\widetilde{N}
) $-type symmetry. However, if the potential generates \textit{%
harmonic-oscillator} states, then these states \emph{will} be
representations of an $SU(\widetilde{N})$-algebra. (See Appendix A.). Since
quarks exhibit $SU(3)$ color-symmetry, this would be a vote for $\widetilde{N%
}=3$ and a symmetrical harmonic-oscillator potential. Because the
coordinates are real, only the (triangular) \textbf{1}, \textbf{3}, \textbf{6%
}, \textbf{10}, . . . \textbf{\ \ }representations would appear, each just
once [5]. The \textbf{3} might correspond to a quark. (The \textbf{1} might
correspond to a ``lepton''. The \textbf{6, 10, . .} would be new species of
Dirac particles.) The spin-degree of freedom might simulate weak isospin.
To accommodate all three generations of quarks and leptons, a higher
symmetry would be required. $SU\left( 4\right) $ might suffice. For four
(real) extra coordinates, the $SU\left( 4\right) $-wavefunctions have the
tetrahedral \textbf{1}, \textbf{4}, \textbf{10}, \textbf{20}, . .
representations, with the \textbf{4} breaking to \textbf{1}$+$\textbf{3},
the \textbf{10} breaking to \textbf{1}$+$\textbf{3}$+$\textbf{6}, \textit{etc%
}. The $SU\left( 4\right) $-\textbf{1} might represent a first-generation
``lepton'', the \textbf{1}$+$\textbf{3 = 4} a second-generation ``lepton''
and a first-generation ``quark'', \textit{etc}. Weight diagrams of the
lowest $SU\left( 4\right) $-multiplets are sketched in Fig. 2.
\begin{figure}[p]
\epsfysize=12cm \centerline{\epsfbox{fig2epsf.ps}}
\caption{Weight diagrams of lowest lying $SU\left( 4\right) \times SU\left(
2\right) $ multiplets of Eqs. 20 and 24. Diagrams plotted \textit{vs}.
principal $SU\left( 4\right) $ quantum number $\widetilde{N}$ and $SU\left(
2\right) $ quantum number $\tilde{m}_{\tilde{s}}$ . ``Quarks'' and
``leptons'' suggested by these quantum numbers are indicated. All members of
a given multiplet are degenerate in mass. However the weight diagrams are
oriented so that heavier particles appear above lighter particles; \emph{i.e.%
}, the $\widetilde{n}_{4}$ (``generation-number'') axis points downward for $%
\tilde{m}_{\tilde{s}}=\frac{1}{2}$ multiplets and upward for $\tilde{m}_{%
\tilde{s}}=-\frac{1}{2}$ multiplets.}
\end{figure}
Could a harmonic-oscillator ``potential'' actually exist in a higher
dimensional flat space? Possibly. Dvali and Shifman have shown that for some
topological defects, in supersymmetric theories at least, such
oscillator-type potentials appear naturally [9].
However a Dirac equation with a harmonic-oscillator potential will not
necessarily generate harmonic-oscillator wavefunctions. A Schr\"{o}dinger
equation or a Klein-Gordon equation will [5], but a Dirac equation might not
because the Dirac operator ``squares'' the potential\footnote{%
Mathematically speaking, a Dirac equation in four Euclidian dimensions
cannot be $SU\left( 4\right) $-symmetric because it incorporates just four $%
\tilde{\gamma}^{\tilde{\mu}}$-matrices, and the corresponding generators $%
\tilde{\gamma}^{\tilde{\mu}}\tilde{\gamma}^{\tilde{\nu}},$ $\mu ,$ $\nu =1,$
$2,\;3,\;4$ can only constitute an $SO\left( 4\right) $-algebra. Adding a
potential will not alter this fact. To generate an $SU\left( 4\right) $
-algebra requires \textit{eight} Dirac matrices [10] such as appear in the
set of $SU\left( 4\right) $-generators $\tilde{\theta}^{\tilde{\mu}^{\dagger
}}\tilde{\theta}^{\tilde{\nu}},$ $\;\mu ,$ $\nu =1,\;2,\;3,\;4,$ where $%
\tilde{\theta}^{\tilde{\mu}^{\dagger }}=\frac{1}{2}\left( -i\tilde{\gamma}^{%
\tilde{\mu}}+\tilde{\gamma}^{\tilde{\mu}+4}\right) $ and $\tilde{\theta}^{%
\tilde{\nu}}=\frac{1}{2}\left( i\tilde{\gamma}^{\tilde{\nu}}+\tilde{\gamma}^{%
\tilde{\nu}+4}\right) .$}. However consider the hydrogen atom in ordinary
space-time$.$ We know that the Coulomb potential in the Dirac equation of
the H-atom reappears as the Coulomb potential in the (Pauli-reduced)
Schr\"{o}dinger equation. This happens because the rest-mass of the electron
is very large compared to the kinetic and potential energies of the electron
and thereby minimizes the symmetry-breaking ``small'' components. Similarly
if we extend Eq. 8 to eight flat dimensions, replace the domain wall with a
symmetrical harmonic-oscillator potential $\mathbf{V}$ in the four extra
dimensions and insert a large ``rest-mass'' $M_{0}$, yielding
\begin{equation}
\left( -i\sum_{\mathbf{\mu }=0}^{7}\mathbf{\gamma }^{\mathbf{\mu }}\partial
_{\mathbf{\mu }}\;\mathbf{+\;V}+M_{0}\right) \mathbf{\psi }=0,
\end{equation}
then $\mathbf{V}$ reappears unsquared in the two-component reduction of the
Dirac equation in the four higher dimensions. This Pauli equation will
generate exact $SU\left( 4\right) $ harmonic-oscillator ``large''
components, and the symmetry-breaking lower components will be small. We
will show this directly.
Let $\mathbf{\gamma }^{\mathbf{\mu }}\mathbf{\gamma }^{\mathbf{\nu }}+%
\mathbf{\gamma }^{\mathbf{\nu }}\mathbf{\gamma }^{\mathbf{\mu }}=2\mathbf{g}
^{\mathbf{\mu \nu }},$\ $\mathbf{\mu ,\nu }=0,$\ $1,$\ $\;7,$\ where the
metric-tensor $\mathbf{g}^{\mathbf{\mu \nu }}$\ =\ diag.$\left( +,-,-,\ldots
,-\right) .$\ A convenient representation of the gamma-matrices is the
``chiral'' set
\begin{equation}
\mathbf{\gamma }^{\mathbf{\mu }}=\left\{
\begin{array}{ll}
\gamma ^{\mu }\otimes \widetilde{\gamma }^{5}, & \mathbf{\mu }=\mu =0,1,2,3,
\\
1\otimes \widetilde{\gamma }^{5}\widetilde{\gamma }^{\widetilde{\mu }}, &
\mathbf{\mu }=\widetilde{\mu }+3=4,5,6,7
\end{array}
\right.
\end{equation}
where $\gamma ^{\mu },$ $\;\mu =0,\;1,\;2,\;3$ is any standard set of Dirac
matrices while
$\widetilde{\gamma }^{\widetilde{j}}=\left(
\begin{array}{ll}
\;\;0 & i\widetilde{\sigma }^{\widetilde{j}} \\
-i\widetilde{\sigma }^{\widetilde{j}} & 0
\end{array}
\right) ,\;\widetilde{j}=1,2,3,$ $\;\;\;\widetilde{\gamma }^{4}=\left(
\begin{array}{ll}
0 & 1 \\
1 & 0
\end{array}
\right) ,$ \ \ and $\qquad \widetilde{\gamma }^{5}=\left(
\begin{array}{ll}
1 & \;\,0 \\
0 & -1
\end{array}
\right) .$ The $\widetilde{\sigma }^{\widetilde{j}}$ are the $2\times 2$
Pauli spin matrices. The potential $\mathbf{V}=1\otimes \tilde{\gamma}^{5}%
\tilde{V}\left( \tilde{x}\right) $, where $\tilde{V}\ $has the
harmonic-oscillator form.
If one multiplies Eq. 11 by $1\otimes \widetilde{\gamma }^{5}$ and sets $%
\mathbf{\psi }=\psi \left( x\right) \otimes \widetilde{\psi }\left( \tilde{x}%
\right) $ , then (11) separates into
\begin{equation}
\left( -i\mathbf{\rlap/\nabla }+M\right) \psi \left( x\right) =0
\end{equation}
and
\begin{equation}
\left( -i\tilde{\rlap/\nabla }+\widetilde{V}+\widetilde{\gamma }%
^{5}M_{0}\right) \widetilde{\psi }=M\widetilde{\psi ,}
\end{equation}
where $\mathbf{\rlap/\nabla }=\sum_{\mu =0}^{3}\gamma ^{\mu }\partial
/\partial x^{\mu }$ and $\tilde{\rlap/\nabla }=\sum_{\widetilde{\mu }=1}^{4}%
\widetilde{\gamma }^{\widetilde{\mu }}\partial /\partial \widetilde{x}^{%
\widetilde{\mu }}$. We have denoted $\left( \mathbf{x}^{4},\mathbf{x}^{5},%
\mathbf{x}^{6},\mathbf{x}^{7}\right) \equiv \left( \widetilde{x}^{1},%
\widetilde{x}^{2},\widetilde{x}^{3},\widetilde{x}^{4}\right) .$ $M$ (not $%
M_{0})$ is taken to be the rest-mass of the Dirac particle in ordinary
space-time$.$ In practice, one solves Eq. 14 for the eigenfunction $%
\widetilde{\psi }$ and eigenmass $M$, then inserts $M$ in Dirac Eq. 13.
The Dirac equation in the extra dimensions may be expanded to
\begin{equation}
\left(
\begin{array}{ll}
\widetilde{V}+M_{0}-M & \qquad \widetilde{D} \\
\qquad \widetilde{D}^{\dagger } & \widetilde{V}-M_{0}-M
\end{array}
\right) \left(
\begin{array}{l}
\widetilde{\phi } \\
\widetilde{\chi }
\end{array}
\right) \;=\;0,
\end{equation}
where
\begin{equation}
\widetilde{D}\;=\;\left(
\begin{array}{ll}
\widetilde{\partial }_{3}-i\widetilde{\partial }_{4}, & \quad \widetilde{
\partial }_{1}-i\widetilde{\partial }_{2} \\
\widetilde{\partial }_{1}+i\widetilde{\partial }_{2}, & -\;\widetilde{
\partial }_{3}-i\widetilde{\partial }_{4}
\end{array}
\right)
\end{equation}
with $\widetilde{\partial }_{\widetilde{\mu }}\equiv \partial /\partial
\widetilde{x}^{\widetilde{\mu }},\;\widetilde{\mu }=1,\;2,\;3,\;4.$ \quad If
$M_{0}\gg \left\langle \widetilde{V}\right\rangle $ and $M\sim M_{0},$ then
\begin{equation}
\widetilde{\chi }\thickapprox \frac{1}{2M_{0}}\widetilde{D}^{\dagger }%
\widetilde{\phi },
\end{equation}
yielding
\begin{equation}
\left( \widetilde{V}+M_{0}-M\right) \widetilde{\phi }\,+\,\widetilde{D}
\left( 2M_{0}\right) ^{-1}\widetilde{D}^{\dagger }\widetilde{\phi }=0.
\end{equation}
$\widetilde{D}\widetilde{D}^{\dagger }=\widetilde{D}^{\dagger }\widetilde{D}
=-\;\tilde{\square}\cdot \widetilde{1}_{2\times 2}$ where $\tilde{\square}$
is the four-dimensional laplacian in $\widetilde{R}^{4},$ so Eq.18 reduces
to
\begin{equation}
\left( -\;\frac{1}{2M_{0}}\tilde{\square}+\widetilde{V}\right) \widetilde{
\phi }=\left( M-M_{0}\right) \widetilde{\phi }.
\end{equation}
Eq.\thinspace 19 is analogous to the Pauli reduction of the time-independent
Dirac equation in ordinary space, with $M-M_{0}$ playing the role of the
kinetic energy. If $\widetilde{V}$ is a symmetrical harmonic-oscillator
potential, then solutions $\tilde{\phi}$ comprise an exact representation of
an $SU\left( 4\right) $-algebra. In particular, if $\widetilde{V}$=$%
\widetilde{\alpha }^{3}\widetilde{X}^{2},$ where $\widetilde{\alpha }$ is an
adjustable real constant and $\widetilde{X}^{2}=\sum_{\widetilde{\mu }
=1}^{4}\left( \widetilde{x}^{\widetilde{\mu }}\right) ^{2},$ then
\begin{equation}
\tilde{\phi}\,=\,\prod_{\tilde{\mu}=1}^{4}\tilde{\phi}_{\tilde{n}_{\tilde{\mu%
}}}\left( \tilde{\beta}\tilde{x}^{\tilde{\mu}}\right) \,\equiv \,\tilde{\phi}
_{\tilde{n}_{1}\tilde{n}_{2}\tilde{n}_{3}\tilde{n}_{4}}\,\equiv \,\tilde{\phi%
}_{\tilde{n}},
\end{equation}
a product of (unnormalized) harmonic-oscillator functions
\begin{equation}
\tilde{\phi}_{\tilde{n}_{\tilde{\mu}}}\left( \tilde{\beta}\tilde{x}^{\tilde{%
\mu}}\right) =H_{\tilde{n}_{\tilde{\mu}}}\left( \tilde{\beta}\tilde{x}^{%
\tilde{\mu}}\right) \exp \left[ -\tfrac{1}{2}\tilde{\beta}^{2}(\tilde{x}^{%
\tilde{\mu}})^{2}\right] ,\;\;\;\;\tilde{\mu}=1,\;\,2,\;3,\;4;
\end{equation}
here $\tilde{\beta}^{4}=2M_{0}\tilde{\alpha}^{3}$ and $H_{\tilde{n}_{\tilde{%
\mu}}}$ is a Hermite polynomial of degree $\tilde{n}_{\tilde{\mu}
}=0,\;1,\;2,\;3,\;.\;.$ The associated eigenmass $M=M_{0}+\left( 2\tilde{N}%
+4\right) (\tilde{\beta}^{2}/2M_{0})\equiv \overline{M}^{\left( \tilde{N}
\right) },$ where
\begin{equation}
\tilde{N}=\tilde{n}_{1}+\tilde{n}_{2}+\tilde{n}_{3}+\tilde{n}
_{4}=0,\;1,\;2,\;3,\;.\;.
\end{equation}
(See Appendix A.)
Actually the solutions to Pauli equation 19 exhibit a higher symmetry than $%
SU\left( 4\right) $ since they also include constant two-component spinors $%
\tilde{\xi}_{\tilde{m}_{\tilde{s}}}={\binom{1}{0}}$ and ${\binom{0}{1}}$
which are representations of $SU\left( 2\right) $.$\ $Thus
\begin{equation}
\tilde{\phi}\,=\,\tilde{\phi}_{\tilde{n}_{1}\tilde{n}_{2}\tilde{n}_{3}\tilde{%
n}_{4}}\tilde{\xi}_{\tilde{m}_{\tilde{s}}}\,\equiv \,\tilde{\phi}_{\tilde{n},%
\tilde{m}_{\tilde{s}}},
\end{equation}
where the $\tilde{\phi}_{\tilde{n},\tilde{m}_{\tilde{s}}}$ are
representations of $SU\left( 4\right) \times SU\left( 2\right) .$
The solutions to the Dirac equation in the extra dimensions, Eq.\thinspace
14, are thus
\begin{equation}
\tilde{\psi}=\left(
\begin{array}{l}
\tilde{\phi} \\
\tilde{\chi}
\end{array}
\right) \approx \left(
\begin{array}{l}
\;\;\;\;\;\;\tilde{\phi}_{\tilde{n},\tilde{m}_{\tilde{s}}} \\
\frac{1}{2M_{0}}\tilde{D}^{\dagger }\tilde{\phi}_{\tilde{n},\tilde{m}_{%
\tilde{s}}}
\end{array}
\right) \equiv \tilde{\psi}_{\tilde{n}_{1}\tilde{n}_{2}\tilde{n}_{3}\tilde{n}
_{4}\tilde{m}_{\tilde{s}}}\equiv \tilde{\psi}_{\tilde{n},\tilde{m}_{\tilde{s}%
}},
\end{equation}
with the same eigenmasses $\overline{M}^{\left( \tilde{N}\right) }.$ \ Since
$\tilde{\chi}\thickapprox \frac{1}{2M_{0}}\tilde{D}^{\dagger }\tilde{\phi}_{%
\tilde{n},\tilde{m}_{\tilde{s}}}\backsim \frac{1}{2M_{0}}\tilde{\beta}\tilde{
\phi}=\left( \frac{\tilde{\alpha}}{2M_{0}}\right) ^{3/4}\tilde{\phi}$, these
lower components vanish in the limit as $\tilde{\alpha}/2M_{0}\rightarrow 0$
. In this limit, solutions $\tilde{\psi}$ will have the same symmetry as the
$\tilde{\phi},$ namely $SU\left( 4\right) \times SU\left( 2\right) .$
How do the $\tilde{\psi}_{\tilde{n},\tilde{m}_{\tilde{s}}}$ suggest quarks
and leptons? Consider the eigenfunctions of the $SU\left( 4\right) \;$%
\textbf{4} with, say, $SU\left( 2\right) $ spin ``up''. The \textbf{4}'s
upper components $\tilde{\phi}$ are $\tilde{x}^{1}\tilde{F}{\binom{1}{0}},\;%
\tilde{x}^{2}\tilde{F}{\binom{1}{0}},\;\tilde{x}^{3}\tilde{F}{\binom{1}{0}},$
and $\tilde{x}^{4}\tilde{F}{\binom{1}{0}},$ where $\tilde{F}\equiv \exp (-%
\frac{1}{2}\tilde{\beta}^{2}\tilde{X}^{2}).$ The triplet-subset $\left(
\tilde{x}^{1},\;\tilde{x}^{2},\;\tilde{x}^{3}\right) \tilde{F}{\binom{1}{0}}$
, say, is an $SU\left( 3\right) $ \textbf{3} and can be coupled to another
\textbf{3 }and an $SU\left( 3\right) $ \textbf{8 (}gluon) to yield an
invariant scalar, as we show in Sec. IV. Thus the\textbf{\ 3 }resembles a
color triplet. In addition, the \textbf{3} with spin ${\binom{1}{0}}$ and
the \textbf{3} from the other \textbf{4 }with spin ${\binom{0}{1}}$
constitute an $SU\left( 2\right) $\textbf{\ 2 }and can be coupled to an
isospin-\textbf{3} intermediate vector-boson to yield another invariant
scalar. Thus the \textbf{2} suggests a weak-isospin doublet. In this way the
$SU\left( 3\right) \times SU\left( 2\right) $ $\left( \mathbf{3},\mathbf{2}
\right) $ resembles a quark.
The remaining wavefunction of the \textbf{4}, $\tilde{x}^{4}\tilde{F}{\binom{
1}{0}}$, will be an $SU\left( 4\right) $ ``color''-singlet, but still a
member of an $SU\left( 2\right) $-doublet. Thus it can be coupled to an
EW-boson, and suggests a lepton.
As noted earlier, weight diagrams of these multiplets are sketched in Fig.
2, for $\tilde{N}=0,\;1,$ and $2$. Dirac particles suggested by these
multiplets are indicated.
Dirac equation 11 predicts only two bound spin states for each ``orbital''
wavefunction $\tilde{\phi}_{\tilde{n}_{1}\tilde{n}_{2}\tilde{n}_{3}\tilde{n}%
_{4}}$, even though the Dirac wavefunction has four components. This is
analogous to the Dirac equation of the hydrogen atom which predicts just two
bound spin states for each orbital wavefunction rather than four; the other
two spin states correspond to positron-proton states and of course are not
bound. If the H-atom's confining potential were a scalar $\left(
-e^{2}/r\right) $ rather than the fourth component of a vector $\left(
-\gamma ^{0}e^{2}/r\right) $, then the positron-proton states would also be
bound, at a negative energy $E\thicksim -m$.
Similarly if the potential \textbf{V} in eight-dimensional equation 11 were $%
\tilde{V}$ rather than $\tilde{\gamma}^{5}\tilde{V},$ then a second spectrum
of confined states would appear with large \emph{lower} components $\tilde{%
\chi}\,\thickapprox \,\tilde{\phi}_{\tilde{n}_{1}\tilde{n}_{2}\tilde{n}_{3}%
\tilde{n}_{4}}\tilde{\xi}_{\tilde{m}_{\tilde{s}}}$ , small upper components $%
\tilde{\phi},$ and masses $M=-\overline{M}^{\left( \tilde{N}\right) }$. The
masses of this spectrum would be identical to those of the positive spectrum
since a negative mass can be interpreted as a positive mass with the
ordinary space-time spinor $\psi $ replaced by $\gamma ^{5}\psi $. This
would yield twice too many ``weak-isospin'' states. Thus the $\tilde{\gamma}%
^{5}\tilde{V}$ structure is indicated.
The potential $\tilde{V}$ generated by a soliton in five-dimensional $%
\mathbf{\phi }^{4}$-theory does not rise to infinity like a pure HO
potential, of course. Instead it reaches an upper limit, as shown in Fig. 1.
Similarly, any potential generated by a soliton in a higher dimensional
space might be expected to rise no higher than some maximum value, call it $%
\tilde{V}_{\infty }$. In the case of Dirac equation 11, if this maximum
height $\tilde{V}_{\infty }$ is $<<M_{0},$ then it is easy to show that the
equation generates confined states with masses $M$ in the range $%
M_{0}<M<M_{0}+V_{\infty }$ and free states (able to propagate anywhere in $%
M^{4}\times \tilde{R}^{4}$) with masses $M$ in the range $M_{0}+\tilde{V}
_{\infty }<M<\infty $. Because Eq. 11 is a relativistic equation, it also
admits free solutions with negative masses in the range $-\infty <M<-M_{0}+%
\tilde{V}_{\infty }$. Again, if the ordinary space-time spinor $\psi $ is
replaced by $\gamma ^{5}\psi ,$ then the masses in this third spectrum
change sign and span the range $M_{0}-\tilde{V}_{\infty }<M<\infty $.
Now this third mass-spectrum of free particles completely overlaps the
discrete spectrum of trapped particles, so is immediately ruled out by
experiment. However if all three spectra are lowered by an amount, say, $%
M_{0},$ so that $M$ is restricted to the three intervals $0<M<\tilde{V}
_{\infty }$, $\tilde{V}_{\infty }<M<\infty ,$ and $-\infty <M<-2M_{0}+\tilde{%
V}_{\infty },\;\;$then the third spectrum is equivalent to $2M_{0}-\tilde{V}
_{\infty }<M<$ $\infty $ and neither it nor the other continuous spectrum
overlaps the discrete spectrum.\thinspace \thinspace Furthermore this lowers
the threshold of discrete particles from $M_{0},$ which is far too massive,
to $0,$ closer to experiment.
Lowering these spectra can be accomplished, admittedly \textit{ad hoc},
simply by subtracting $M_{0}$ from $\tilde{V}$. Thus we will set
\begin{equation}
\mathbf{V}=1\otimes \tilde{\gamma}^{5}\left( \tilde{V}-M_{0}\right) ,
\end{equation}
where $\tilde{V}$ is a HO-like potential that rises only to $\tilde{V}
_{\infty }$, analogous to the $\tilde{V}$ in Fig. 1. If lim$_{\tilde{X}
\rightarrow 0}\tilde{V}=\tilde{\alpha}^{3}\tilde{X}^{2}$ and $\left( 2\tilde{
N}+4\right) \tilde{\beta}^{2}/2M_{0}<<\tilde{V}_{\infty }$, then
\begin{equation}
M\thickapprox \left( 2\tilde{N}+4\right) (\tilde{\beta}^{2}/2M_{0})=\left( 2%
\tilde{N}+4\right) (\tilde{\alpha}^{3}/2M_{0})^{1/2}\equiv M^{\left( \tilde{N%
}\right) },
\end{equation}
$\tilde{N}=0,\;1,\;2,\;.\;.\;.$
The possibility of continuous mass spectra raises an interesting question.
If we really are trapped in a topological or non-topological defect in a
higher-dimensional flat space, then are there particles freely ranging in
this extended space that could invade our local territory? Could they
sometimes appear even at energies above the Greisen-Zatsepin-Kus'min cutoff $%
[11]$ $(\thicksim 5\cdot 10^{19}$ eV)? And might we be able to propel Dirac
particles from our local space into that continuum? The threshold for
ejecting particles into the conjectured continuum may lie just above the
capability of present-day machines. In that case, there might be events such
as $e^{+}+e^{-}\rightarrow $ $2\;$free\ Dirac\ particles [1], appearing to
be $e^{+}+e^{-}\rightarrow $ nothing.
An interesting question is also posed by the discrete spectrum. If quarks
and leptons really do occur in $SU\left( 4\right) $ multiplets, then $%
SU\left( 3\right) $ \textbf{6}s should accompany the charm and strange
quarks in the $SU\left( 4\right) $ \textbf{10}s (see Fig. 2), and $SU\left(
3\right) $ \textbf{6}s and \textbf{10}s should accompany the top and bottom
quarks in $SU\left( 4\right) $ \textbf{20}s (not shown). Such particles have
never been seen$\footnote{%
Sextet Dirac particles have been conjectured for a long time. See, \textit{\
e.g.}, Ref. [12] and references cited therein.}$. If they do exist, then
they apparently do not interact with either the electroweak bosons or
gluons. This should make them good candidates for dark matter.
The discrete spectrum also suggests fourth-generation leptons, call them $%
\nu _{4}$ and $l_{4}.$ These particles would accompany the top and bottom
quarks in $SU\left( 4\right) $ \textbf{20}s. Such particles have never been
seen either. Present-day experimental limits [11] would place them above
about $45$ GeV.
Of course, just as in nuclear physics, not every member of a multiplet need
be bound. This might explain the absence of at least some of the \textbf{6}s
and \textbf{10}s and leptons.
Another interesting question has to do with the size of the bound states in
the extra dimensions. Their widths can be easily calculated if the confining
``potential'' is taken to be a harmonic oscillator. If we adopt as a measure
of this width (squared)
\begin{equation}
\int \tilde{\psi}_{\widetilde{n}}^{\dagger }\tilde{X}^{2}\tilde{\psi}_{%
\widetilde{n}}d^{4}\tilde{x}\approx \int \tilde{\phi}_{\widetilde{n}%
}^{\dagger }\tilde{X}^{2}\tilde{\phi}_{\widetilde{n}}d^{4}\tilde{x}\equiv
\tilde{X}_{rms}^{2}
\end{equation}
[recall that $\widetilde{X}^{2}=\sum_{\widetilde{\mu }=1}^{4}\left(
\widetilde{x}^{\widetilde{\mu }}\right) ^{2}$]$,$ then one can readily show
that $\tilde{X}_{rms}^{2}=\left( \tilde{N}+2\right) /\tilde{\beta}^{2}.$
Folding in Eq. 26 and restoring $\mathbf{\rlap/}h$ and $c$ yield
\begin{equation}
\tilde{X}_{\widetilde{N}}=\left( \tilde{N}+2\right) \left( M^{\left(
\widetilde{N}\right) }M_{0}\right) ^{-\frac{1}{2}}\mathbf{\rlap/}h/c\equiv
\tilde{X}_{rms}.
\end{equation}
As for numbers, if we choose $M_{0}=100\;$TeV and adjust $\tilde{\alpha}$ to
fit an up-quark mass of $6$ MeV, then $\tilde{X}_{rms}=0.024$\ F (with $%
\tilde{\alpha}=585$ MeV). If we set $M_{0}=10^{4}\;$TeV, then $\tilde{X}%
_{rms}=0.0024$\ F (with $\tilde{\alpha}=2710$ MeV).
Now according to the Particle Data Group [11], leptons and quarks have radii
no greater than the order of $10^{-4}$ F. However these are radii in \emph{%
ordinary} space-time. They can be measured, \textit{e.g.}, by probing the
Dirac particle with photon-wavefunctions $\exp \left( -ip\cdot x\right) $
over a range of momenta $\overrightarrow{p}$ and taking the Fourier
transform. To measure the size of a Dirac particle in both ordinary space
and extended space, one could figuratively probe the particle with
``photon'' wavefunctions $\exp \left( -ip\cdot x+i\tilde{p}\cdot \tilde{x}
\right) $ where now the higher dimensional ``momentum'' $\tilde{p}$ would
also range over some interval. However these would be the wavefunctions of
probes free to propagate in all of the dimensions, higher and lower. Even if
such particles were to exist, we would only have access to the photons
confined in $M^{4}.$ These ordinary photons have just a single wavefunction
in the higher dimensions (more kinds of wavefunctions would mean more kinds
of photons), so they always weight the wavefunction of the Dirac particle
the same way in the higher dimensions and tell us nothing about its
higher-dimensional structure. If this is typical of all types of
higher-dimensional reactions measurable by us, then the size of the Dirac
particles in the extra dimensions is not subject to the Particle Data Group
limits.
\section{Interaction of Dirac particles with gauge fields}
\subsection{QCD interaction}
In this section we will indicate how the ``quark'' solutions of Eq. 11 might
be combined with a ``gluon'' gauge field to reproduce the QCD interaction
lagrangian in the limit as $V_{\infty }\rightarrow \infty $. Consider the
eight-dimensional interaction lagrangian
\begin{equation}
\mathcal{L}_{int}=\tfrac{1}{\sqrt{2}}g_{s}\sum_{\mu =0}^{3}\sum_{\tilde{i},%
\tilde{j}=1}^{3}\overline{\mathbf{\psi }}\gamma ^{\mu }\tilde{E}_{\tilde{i}%
\tilde{j}}\mathbf{G}_{\mu }^{\tilde{i}\tilde{j}}\mathbf{\psi ,}
\end{equation}
where $\mathbf{G}_{\mu }^{\tilde{i}\tilde{j}}$ is the ``gauge field'' and
the $\tilde{E}_{\tilde{i}\tilde{j}}$ are the $SU\left( 3\right) $-generators
defined by Eq. 47 in Appendix A. It may not be unreasonable to suppose that $%
\mathbf{G}_{\mu }^{\tilde{i}\tilde{j}}=G_{\mu }(x)\tilde{G}^{\tilde{i}\tilde{
j}}(\tilde{x})$, in which case ordinary space-time quantities such as
energy-momentum, angular momentum, \textit{etc}. are conserved. In this
case, $\mathcal{L}_{int}$ will be a sum over terms of the sort
\begin{equation}
\tfrac{1}{\sqrt{2}}g_{s}\overline{\psi }_{f}\left( x\right) \gamma ^{\mu
}G_{s,\mu }\left( x\right) \psi _{i}\left( x\right) \cdot \overline{\tilde{
\psi}}_{\tilde{f}}(\tilde{x})\tilde{E}_{\tilde{i}\tilde{j}}\tilde{G}_{\tilde{%
s}}^{\tilde{i}\tilde{j}}(\tilde{x})\tilde{\psi}_{\tilde{i}}(\tilde{x}),
\end{equation}
where Dirac and Bose states are identified by $f$, $i$, and $s$ in $M^{4}$
and $\tilde{f}$, $\tilde{i}$, and $\tilde{s}$ in $\tilde{R}^{4}$. ($%
\overline{\tilde{\psi}}=\tilde{\psi}^{\dagger }\tilde{\gamma}^{5}.)$ In the
case of the bound-state solutions $\tilde{\psi}_{\widetilde{n}_{1}\widetilde{%
n}_{2}\widetilde{n}_{3}\widetilde{n}_{4}\widetilde{m}_{\tilde{s}}},$ the
field $\tilde{G}_{\tilde{s}}^{\tilde{i}\tilde{j}}(\tilde{x})$ can cause
transitions to different ``families'' ($\widetilde{n}_{4})$, to different $%
SU\left( 3\right) $-multiplets $(\widetilde{N}^{\prime }=\widetilde{n}_{1}+%
\widetilde{n}_{2}+\widetilde{n}_{3})$ (\emph{e.g., }quark$\rightarrow $
lepton), and to the opposite ``weak-isospin'' state ($\widetilde{m}_{\tilde{s%
}}$), all of which are inconsistent with QCD. However if $\tilde{G}_{\tilde{s%
}}^{\tilde{i}\tilde{j}}(\tilde{x})$ is essentially constant over the range
of integration of $\overline{\tilde{\psi}}_{\tilde{f}}(\tilde{x})\tilde{E}_{%
\tilde{i}\tilde{j}}\tilde{G}_{\tilde{s}}^{\tilde{i}\tilde{j}}(\tilde{x})%
\tilde{\psi}_{\tilde{i}}(\tilde{x})$ in $\tilde{R}^{4}$, then none of these
transitions can take place due to the orthogonality of the
harmonic-oscillator functions, at least in the limit $\tilde{\alpha}
/2M_{0}\rightarrow 0$ where $\tilde{\psi}\rightarrow \left( \tilde{\phi}_{%
\widetilde{n}_{1}\widetilde{n}_{2}\widetilde{n}_{3}\widetilde{n}_{4,}%
\widetilde{m}_{\tilde{s}}}(\tilde{x}),\;0\right) ^{T}$. An example of a
field that would be essentially constant over the range of integration is $%
\tilde{G}_{\tilde{s}}^{\tilde{i}\tilde{j}}(\tilde{x})=\widehat{G}_{\tilde{s}%
}^{\tilde{i}\tilde{j}}\exp \left( -\frac{1}{2}\tilde{\gamma}^{2}\tilde{X}
^{2}\right) $, where the range $\tilde{\gamma}^{-1}$ of the gluon field is
much greater than the range $\tilde{\beta}^{-1}$ of the Dirac-particle
fields $\left( \tilde{\gamma}^{-1}>>\tilde{\beta}^{-1}\right) $; $\widehat{G}
_{\tilde{s}}^{\tilde{i}\tilde{j}}$ is a constant which identifies the
particular gluon field $\tilde{s}$.
Lagrangian 30 is still not consistent with QCD since $\tilde{\psi}$ can be
any $SU\left( 3\right) $-multiplet: \textbf{1}, \textbf{3}, \textbf{6},
\textbf{10}, . . . However if we limit the $SU\left( 3\right) $-states to
just \textbf{3}s ($\widetilde{n}_{1}+\widetilde{n}_{2}+\widetilde{n}_{3}=1$
``quarks'') and \textbf{1}s ($\widetilde{n}_{1}+\widetilde{n}_{2}+\widetilde{%
n}_{3}=0$ ``leptons''), then the lagrangian is equivalent to the QCD
lagrangian\footnote{%
If by no other means, one can project out \textbf{6}s (at least at tree
level) by inserting the operator $\left( 2-\sum_{\tilde{k}=1}^{3}\widetilde{a%
}_{\tilde{k}}^{\dagger }\widetilde{a}_{\tilde{k}}\right) $ before or after $%
\tilde{E}_{\tilde{i}\tilde{j}}$ in lagrangian 29. Higher multiplets, if they
are exist (\emph{i.e.}, are bound), can be projected out by similar
operators. \textbf{1}s make no contribution since $\tilde{E}_{\tilde{i}%
\tilde{j}}\tilde{\phi}_{000}=0.$}, as we will now show. One can integrate
over $\tilde{R}^{4}$ and replace the $\tilde{\phi}_{\widetilde{n}_{1}%
\widetilde{n}_{2}\widetilde{n}_{3}\widetilde{n}_{4,}\widetilde{m}_{\tilde{s}%
}}(\tilde{x})$ by unit column vectors $\widehat{\phi }_{\widetilde{n}_{1}%
\widetilde{n}_{2}\widetilde{n}_{3}\widetilde{n}_{4,}\widetilde{m}_{\tilde{s}%
}}=\widehat{\phi }_{\widetilde{n}_{1}\widetilde{n}_{2}\widetilde{n}%
_{3}}\otimes \widehat{\phi }_{\widetilde{n}_{4}}\otimes \tilde{\xi}_{\tilde{m%
}_{\tilde{s}}}$, where the $\widehat{\phi }_{\widetilde{n}_{1}\widetilde{n}%
_{2}\widetilde{n}_{3}}\equiv \widehat{\phi }_{color}$ comprise the ``color''
states [ $\widehat{\phi }_{100}=(1,0,0)^{T},$ $\widehat{\phi }%
_{010}=(0,1,0)^{T},$ and $\widehat{\phi }_{001}=(0,0,1)^{T}$], $\widehat{%
\phi }_{\widetilde{n}_{4}}$ is a column vector with an element for each
``generation'', and $\tilde{\xi}_{\tilde{m}_{\tilde{s}}}=(1,0)^{T}$ or $%
(0,1)^{T}$, the previously defined ``weak-isospin'' spinor. The operators $%
\tilde{E}_{\tilde{i}\tilde{j}}(\tilde{x},$ $\partial \tilde{x})$ can
similarly be replaced by $3\times 3$ matrices $\widehat{E}_{\tilde{i}\tilde{j%
}},$ where
\begin{equation}
\widehat{E}_{11}=\left(
\begin{array}{ccc}
\frac{2}{3} & 0 & 0 \\
0 & -\frac{1}{3} & 0 \\
0 & 0 & -\frac{1}{3}
\end{array}
\right) ,\;\widehat{E}_{12}=\left(
\begin{array}{lll}
0 & 1 & 0 \\
0 & 0 & 0 \\
0 & 0 & 0
\end{array}
\right) ,\;\widehat{E}_{13}=\left(
\begin{array}{lll}
0 & 0 & 1 \\
0 & 0 & 0 \\
0 & 0 & 0
\end{array}
\right)
\end{equation}
and similarly for $\widehat{E}_{2\tilde{j}}$ and $\widehat{E}_{3\tilde{j}},\;%
\tilde{j}=1,\;2,\;3$. The interaction lagrangian now takes the form
\begin{equation}
\widehat{\mathcal{L}}_{int}=\tfrac{1}{\sqrt{2}}g_{s}\sum_{\mu =0}^{3}\sum_{%
\tilde{i},\tilde{j}=1}^{3}\overline{\widehat{\mathbf{\psi }}}\gamma ^{\mu }%
\widehat{E}_{\tilde{i}\tilde{j}}\widehat{\mathbf{G}}_{\mu }^{\tilde{i}\tilde{
j}}\widehat{\mathbf{\psi }},
\end{equation}
where $\widehat{\mathbf{\psi }}$ is a sum over states $\psi \left( x\right)
\otimes \widehat{\phi }_{color}\otimes \widehat{\phi }_{\widetilde{n}
_{4}}\otimes \tilde{\xi}_{\tilde{m}_{\tilde{s}}}$ and $\widehat{\mathbf{G}}%
_{\mu }^{\tilde{i}\tilde{j}}$ is a sum over states $G_{\mu }\left( x\right)
\widehat{G}^{\tilde{i}\tilde{j}}$. This lagrangian equals the QCD lagrangian
[11]
\begin{equation}
\mathcal{L}_{int}^{\scriptstyle QCD}=\tfrac{1}{2}g_{s}\sum_{\mu
=0}^{3}\sum_{a=1}^{8}\overline{\widehat{\mathbf{\psi }}}\gamma ^{\mu
}\lambda ^{a}\mathbf{G}_{\mu }^{a}\widehat{\mathbf{\psi }},
\end{equation}
where the $\lambda ^{a}$ are the Gell-Mann $3\times 3\;SU\left( 3\right) $
matrices, if the QCD gauge fields
\[
\mathbf{G}_{\mu }^{1}=\tfrac{1}{\sqrt{2}}\left( \widehat{\mathbf{G}}_{\mu
}^{12}+\widehat{\mathbf{G}}_{\mu }^{21}\right) ,\ \;\mathbf{G}_{\mu }^{2}=%
\tfrac{i}{\sqrt{2}}\left( \widehat{\mathbf{G}}_{\mu }^{12}-\widehat{\mathbf{G%
}}_{\mu }^{21}\right) ,\;\;
\]
\[
\mathbf{G}_{\mu }^{3}=\tfrac{1}{\sqrt{2}}\left( \widehat{\mathbf{G}}_{\mu
}^{11}-\widehat{\mathbf{G}}_{\mu }^{22}\right) ,\;\;\mathbf{G}_{\mu }^{4}=%
\tfrac{1}{\sqrt{2}}\left( \widehat{\mathbf{G}}_{\mu }^{13}+\widehat{\mathbf{G%
}}_{\mu }^{31}\right) ,\;\;
\]
\[
\mathbf{G}_{\mu }^{5}=\tfrac{i}{\sqrt{2}}\left( \widehat{\mathbf{G}}_{\mu
}^{13}-\widehat{\mathbf{G}}_{\mu }^{31}\right) ,\;\;\mathbf{G}_{\mu }^{6}=%
\tfrac{1}{\sqrt{2}}\left( \widehat{\mathbf{G}}_{\mu }^{23}+\widehat{\mathbf{G%
}}_{\mu }^{32}\right) ,\;\;
\]
\begin{equation}
\mathbf{G}_{\mu }^{7}=\tfrac{i}{\sqrt{2}}\left( \widehat{\mathbf{G}}_{\mu
}^{23}-\widehat{\mathbf{G}}_{\mu }^{32}\right) ,\ \ \mathbf{G}_{\mu }^{8}=%
\tfrac{1}{\sqrt{6}}\left( \widehat{\mathbf{G}}_{\mu }^{11}+\widehat{\mathbf{G%
}}_{\mu }^{22}-2\widehat{\mathbf{G}}_{\mu }^{33}\right) .\bigskip
\end{equation}
\subsection{Electroweak interaction}
We can also construct a lagrangian which reduces to the electroweak
interaction, provided the latter is of the ``left-right symmetric'' variety
[7]. Consider the eight-dimensional lagrangian
\[
\mathcal{L}_{int}=\tfrac{1}{2}\sum_{\mu =0}^{3}\sum_{\tilde{j}=1}^{3}\left[
g_{L}\overline{\mathbf{\psi }}\gamma ^{\mu }\tfrac{1}{2}\left( 1-\gamma
^{5}\right) \tilde{\Sigma }^{\tilde{j}}\mathbf{W}_{L,\mu }^{\tilde{j}}%
\mathbf{\psi }\right. \mathbf{\,}
\]
\begin{equation}
\left. \mathbf{+\,}~g_{R}\overline{\mathbf{\psi }}\gamma ^{\mu }\tfrac{1}{2}%
\left( 1+\gamma ^{5}\right) \tilde{\Sigma }^{\tilde{j}}\mathbf{W}_{R,\mu }^{%
\tilde{j}}\mathbf{\psi }\right] \mathbf{,}
\end{equation}
where $\mathbf{W}_{L,\mu }^{\tilde{j}}$ and $\mathbf{W}_{R,\mu }^{\tilde{j}}$
are ``electro-weak'' gauge-field triplets and $\tilde{\Sigma }^{\tilde{j}%
}=\left(
\begin{array}{ll}
\tilde{\sigma}^{\tilde{j}} & 0 \\
0 & \tilde{\sigma}^{\tilde{j}}
\end{array}
\right) ,$ $\tilde{j}=1,\;2,\;3$. If we again assume that gauge fields can
be factored, \textit{i.e.}, $\mathbf{W}_{L,\mu }^{\tilde{j}}=W_{L,\mu
}\left( x\right) \tilde{W}^{\tilde{j}}(\tilde{x})$ and $\mathbf{W}_{R,\mu }^{%
\tilde{j}}=W_{R,\mu }\left( x\right) \tilde{W}^{\tilde{j}}(\tilde{x}),$ then
Minkowski--space quantities will be conserved. If in addition we
parameterize $\tilde{W}^{\tilde{j}}(\tilde{x})=\widehat{W}^{\tilde{j}}\exp
\left( -\frac{1}{2}\tilde{\gamma}^{2}\tilde{X}^{2}\right) $ where $\widehat{W%
}^{\tilde{j}}$ is a constant identifying the boson, and again assume that $%
\tilde{\gamma}^{-1}>>\tilde{\beta}^{-1}$, then in the limit $\tilde{\alpha}%
/2M_{0}\rightarrow 0$ where orthogonality holds, Dirac particles may not
make transitions to different ``generations'', to different $SU\left(
3\right) $-multiplets, or to different states within an $SU\left( 3\right) $
-multiplet. However transitions to opposite ``weak-isospin'' states may take
place due to the $\tilde{\sigma}^{\tilde{j}}$-operator. If again we limit
Dirac states to $SU\left( 3\right) $ \textbf{1}s and \textbf{3}s, then
integrating lagrangian 35 over $\tilde{R}^{4}$ yields
\[
\widehat{\mathcal{L}}_{int}=\tfrac{1}{2}\sum_{\mu =0}^{3}\sum_{\tilde{j}%
=1}^{3}\left[ g_{L}\overline{\widehat{\mathbf{\psi }}}\gamma ^{\mu }\tfrac{1%
}{2}\left( 1-\gamma ^{5}\right) W_{L,\mu }\left( x\right) \widetilde{\sigma }%
^{\tilde{j}}\widehat{W}^{\tilde{j}}\widehat{\mathbf{\psi }}\right.
\]
\begin{equation}
\left. \mathbf{+\,\,}g_{R}\overline{\widehat{\mathbf{\psi }}}\gamma ^{\mu }%
\tfrac{1}{2}\left( 1+\gamma ^{5}\right) W_{R,\mu }\left( x\right) \widetilde{%
\sigma }^{\tilde{j}}\widehat{W}^{\tilde{j}}\widehat{\mathbf{\psi }}\right] ,
\end{equation}
where operator $\widehat{\mathbf{\psi }}$ is a sum over states $\psi \left(
x\right) \otimes \widehat{\phi }_{\widetilde{n}_{1}\widetilde{n}_{2}%
\widetilde{n}_{3}}\otimes \widehat{\phi }_{\widetilde{n}_{4}}\otimes \tilde{%
\xi}_{\widetilde{m}_{\tilde{s}}}$with $\widetilde{n}_{1}+\widetilde{n}_{2}+%
\widetilde{n}_{3}=0$ or $1.$ If we denote $\overrightarrow{\tilde{\sigma}}%
\equiv \overrightarrow{\tau }$, $W_{L,\mu }\left( x\right) \widehat{W}^{%
\tilde{j}}\equiv \mathbf{W}_{L,\mu }^{\tilde{j}}\left( x\right) ,$ and $%
W_{R,\mu }\left( x\right) \widehat{W}^{\tilde{j}}\equiv \mathbf{W}_{R,\mu }^{%
\tilde{j}}\left( x\right) ,\;\tilde{j}=1,\;2,\;3,$ then lagrangian 36 takes
the form\pagebreak
\[
\widehat{\mathcal{L}}_{int}=\tfrac{1}{2}g_{L}\overline{\widehat{\mathbf{\psi
}}}\gamma ^{\mu }\tfrac{1}{2}\left( 1-\gamma ^{5}\right) \overrightarrow{%
\tau }\cdot \overrightarrow{\mathbf{W}}_{L,\mu }\left( x\right) \widehat{%
\mathbf{\psi }}\mathbf{\;}
\]
\begin{equation}
\mathbf{+\;}\tfrac{1}{2}g_{R}\overline{\widehat{\mathbf{\psi }}}\gamma ^{\mu
}\tfrac{1}{2}\left( 1+\gamma ^{5}\right) \overrightarrow{\tau }\cdot
\overrightarrow{\mathbf{W}}_{R,\mu }\left( x\right) \widehat{\mathbf{\psi }}.
\end{equation}
This lagrangian forbids family-changing transitions, quark$\leftrightarrow $
lepton transitions, and, in the case of quarks, color-changing transitions.
If a fourth, weak-isospin-zero gauge field $B^{\mu }$ is introduced with
appropriate, distinct couplings to quark and lepton fields, and CKM mixing
is accommodated, then the interaction lagrangian becomes a
left-right-symmetric version of the EW interaction in the Standard Model.
\section{Discussion}
We have considered the possibility that quarks and leptons live in a flat
higher dimensional space, confined to a small volume in the extra
dimensions, perhaps by a soliton. If the confinement is due to a
harmonic-oscillator ``potential'' which only rises to a certain height, plus
a large constant $M_{0}$, then four extra dimensions suffice to generate
harmonic-oscillator Dirac-particle states which bear some resemblance to the
quarks and leptons of the Standard Model. The symmetry group of the states
in the higher dimensions is just $SO\left( 4\right) $, but thanks to the
degenerate levels of the harmonic-oscillator potential, the $SO\left(
4\right) $ symmetry is enlarged to an approximate $SU\left( 4\right) $
symmetry. Furthermore, because the ``spin'' states are decoupled from the
``orbital'' states, an additional approximate $SU\left( 2\right) $ symmetry
occurs in these same four extra dimensions. Thus the HO potential generates
an approximate $SU\left( 4\right) \times SU\left( 2\right) $ symmetry.
In an $SU\left( 4\right) =SU\left( 3\right) \times U\left( 1\right) $
decomposition, the $SU\left( 3\right) $ representations \textbf{1, 3},
\textbf{6}, . . resemble leptons, quarks, and new kinds of Dirac particles,
respectively; the $U\left( 1\right) $ representations introduce replications
which might be identified with generations. The $SU\left( 2\right) $
symmetry is represented by doublets reduced from Dirac spinors and is
similar to weak isospin.
We have shown how the Dirac particles might be coupled to ``gluons'' to
reproduce the QCD interaction, and to ``EW bosons'' to reproduce a
left-right symmetric extension of the electroweak interaction. Because the $%
SU\left( 4\right) $ symmetry is realized in just four dimensions instead of
eight, only real solutions corresponding to tetrahedral weight diagrams are
generated. But these are the only multiplets needed to represent the quarks
and leptons of the Standard Model.
The $SU\left( 4\right) $ symmetry is not exact, primarily because the
harmonic-oscillator well rises only a finite amount. On the other hand, this
finite rise limits the number of quarks and leptons in a natural way. (The
symmetry in the lower generations would become more exact should the well
rise higher to accommodate additional generations.)
Also the masses predicted by the harmonic-oscillator well rise only linearly
with generation number, whereas in nature the masses seem to rise
exponentially. However the particles' masses might be augmented by a
Higgs-Yukawa mechanism as in the Standard Model.
If a simple potential in four flat extra dimensions can yield Dirac bound
states similar in many ways to those of quarks and leptons, then it may be
worthwhile to search for an underlying physical mechanism, perhaps an
instanton arising from a generalization of the Rubakov-\-Shaposhnikov [1] or
Kaplan [2] lagrangian to eight dimensions. Hopefully such a lagrangian will
not generate the $SU\left( 3\right) $ \textbf{6}s and \textbf{10}s which
occur in the higher $SU\left( 4\right) $ multiplets predicted by our present
model. On the other hand, if the \textbf{6}s and \textbf{10}s were decoupled
from the gauge bosons, then they would be interesting candidates for dark
matter.
\section*{Acknowledgments}
I would like to thank George Sudarshan, Dick Arnowitt, Chris Pope, Mike
Duff, Ergin Sezgin and Rabindra Mohapatra for helpful discussions.
\section*{Note added in proof}
Recently there has been considerable interest in the possibility that the
extra dimensions, while not flat or infinite in extent, may still be of
millimeter size. Akrani-Hamed, Dimopoulos, Dvali, and Antoniadis and others
have proposed models of this sort. See, \emph{e. g.}, references [14] --
[16]. These models have passed a gauntlet of tests, but many remain. See,
\emph{e. g.}, reference [17].
\section*{References}
1. V.~A.~Rubakov and M.~E.~Shaposhnikov, ``Do we live inside a domain
wall?'', Phys. Lett. B \textbf{125,} 136 (1983).
2. D. B. Kaplan, ``A method for simulating chiral fermions on the lattice'',
Phys. Lett. B \textbf{288,} 342 (1992).
3. H. B. Nielsen and M. Ninomiya, ``A no-go theorem for regularizing chiral
fermions'', Phys. Lett. B \textbf{105,} 219 (1981); L. H. Karsten and
J.~Smit, ``Lattice fermions: species doubling, chiral invariance and the
triangle anomaly'', Nucl. Phys. B \textbf{183,} 103 (1981).
4. K. Jansen, ``Domain wall fermions and chiral gauge theories'', Phys. Rep.
\textbf{273, }1\textbf{\ }(1996).
5. R.~A.~Bryan, ``Isoscalar quarks and leptons in an eight-dimensional
space'', Phys. Rev. D \textbf{34,} 1184 (1986).
6. G. Dvali and M. Shifman, ``Domain walls in strongly coupled theories'',
Phys. Lett. B \textbf{396,} 64 (1997); hep-th/9612128.
7. \textit{e.g.}, R. N. Mohapatra and G. Senjanovi\'{c}, ``Neutrino masses
and mixings in gauge models with spontaneous parity violation'',
Phys.~Rev.~D \textbf{23,} 165 (1981).
8. R. Rajaraman, \emph{Solitons and instantons} (Amsterdam, North-Holland,
1982) p.~140.
9. G. Dvali and M. Shifman, ``Dynamical compactification as a mechanism of
spontaneous supersymmetry breaking'', Nucl$.$ Phys. B \textbf{504,} 127
(1997); hep-th/9611213.
10. \textit{e.g.}, R. A. Bryan and M. M. Davenport, ``Spinorial
representations of SU(3) from a factored harmonic-oscillator equation'', J.
Phys. A \textbf{29,} 3129 (1996).
11. R. M. Barnett \emph{et al.} (Particle Data Group), ``Review of particle
physics'', Phys. Rev. D \textbf{54,} 1 (1996); K. Greisen, Phys. Rev. Lett.
\textbf{16}, 748 (1966); G. T. Zatsepin and V. A. Kuz'min, Pis'ma Zh.
\'{E}ksp. Teor. Fiz. \textbf{4}, 114 (1966) [JETP Lett. \textbf{4}, 78
(1966].
12. P. H. Frampton and S. L. Glashow, ``Unifiable chiral color with natural
Glashow-Iliopoulos-Maiani mechanism'', Phys. Rev. Letters \textbf{58,} 2168
(1987); T. E. Clark, C. N. Leung, S. T. Love, and J. L. Rosner, ``Sextet
quarks and light pseudoscalars'', Phys. Letters B \textbf{177,} 413 (1986).
13. H.~J.~Lipkin, \emph{Lie Groups for Pedestrians, 2nd edn} (Amsterdam,
North-Holland, 1966) p.~33.
14. N. Arkani--Hamed, S. Dimopoulos, and G. Dvali, ``The hierarchy problem
and new dimensions at a millimeter'', Physics Letters B \textbf{429}, 263
(1998); hep-ph/9803315.
15. I. Antoniadis, N. Arkani--Hamed, S. Dimopoulos, and G. Dvali, ``New
dimensions at a millimeter to a fermi and superstrings at a TeV'', Physics
Letters B \textbf{436}, 257 (1998); hep-ph/9804398.
16. N. Arkani--Hamed, S. Dimopoulos, and G. Dvali, ``Phenomenology,
astrophysics and cosmology of theories with sub-millimeter dimensions and
TeV scale quantum gravity'', Physical Review D \textbf{59}, 086004 (1999);
hep-ph/9807344.
17. T. Banks, M. Dine, and A. E. Nelson, ``Constraints on theories with
large extra dimensions'', hep-th/9903019 v3.
\section*{Appendix A: Harmonic-oscillator symmetry groups}
Consider the differential equation
\begin{equation}
\left( \widetilde{\square }-\tilde{\beta}^{4}\tilde{X}^{2}+m^{2}\right)
\widetilde{\phi }=0
\end{equation}
in four Euclidean extra dimensions $\tilde{R}^{4}\left( \tilde{x}^{1},\;%
\tilde{x}^{2},\;\tilde{x}^{3},\;\tilde{x}^{4}\right) ,$ with $\tilde{\beta}$
a real constant, $\widetilde{\square }$ $=\sum_{\widetilde{\mu }
=1}^{4}\partial ^{2}/\partial \left( \widetilde{x}^{\tilde{\mu}}\right) ^{2}$
and $\tilde{X}^{2}$ $=\sum_{\widetilde{\mu }=1}^{4}\left( \widetilde{x}^{%
\tilde{\mu}}\right) ^{2}$. This equation has (unnormalized) solutions
\begin{equation}
\widetilde{\phi }=\widetilde{\phi }_{\widetilde{n}_{1}}\left( \tilde{\beta}%
\widetilde{x}^{1}\right) \widetilde{\phi }_{\widetilde{n}_{2}}\left( \tilde{
\beta}\widetilde{x}^{2}\right) \widetilde{\phi }_{\widetilde{n}_{3}}\left(
\tilde{\beta}\widetilde{x}^{3}\right) \widetilde{\phi }_{\widetilde{n}%
_{4}}\left( \tilde{\beta}\widetilde{x}^{4}\right) \equiv \widetilde{\phi }_{%
\widetilde{n}_{1}\widetilde{n}_{2}\widetilde{n}_{3}\widetilde{n}_{4}}
\end{equation}
with eigenvalues
\begin{equation}
m^{2}=\left( 2\widetilde{N}+4\right) \tilde{\beta}^{2},
\end{equation}
where
\begin{equation}
\widetilde{\phi }_{\widetilde{n}_{\widetilde{\mu }}}\left( \tilde{\beta}%
\widetilde{x}^{\widetilde{\mu }}\right) =H_{\widetilde{n}_{\widetilde{\mu }%
}}\left( \widetilde{\beta }\widetilde{x}^{\widetilde{\mu }}\right) \exp
\left[ -\tfrac{1}{2}\tilde{\beta}^{2}(\tilde{x}^{\tilde{\mu}})^{2}\right]
,\;\;\;\;\widetilde{\mu }=1,\ 2,\ 3,\ 4.
\end{equation}
Here $H_{\widetilde{n}_{\widetilde{\mu }}}$ is a Hermite polynomial of
degree $\widetilde{n}_{\widetilde{\mu }}=0,$\ $1,\;2,\;.\;.$, and$\;%
\widetilde{N}=\sum_{\widetilde{\mu }=1}^{4}\widetilde{n}_{\tilde{\mu}}=0,\
1,\;2,\;.\;$.
Solutions of a given $\widetilde{N}\,\,\,$form multiplets of common mass and
can be transmuted one into another by the generators
\begin{equation}
\widetilde{a}_{\tilde{\mu}}^{\dagger }\widetilde{a}_{\tilde{\nu}}-\tfrac{1}{4%
}\delta _{\tilde{\mu}\tilde{\nu}}\sum_{\widetilde{\rho }=1}^{4}\widetilde{a}%
_{\tilde{\rho}}^{\dagger }\widetilde{a}_{\tilde{\rho}}\equiv \tilde{E}_{%
\tilde{\mu}\tilde{\nu}},\;\;\;\mu ,\;\nu =1,\;2,\;3,\;4,
\end{equation}
where
\begin{equation}
\widetilde{a}_{\tilde{\mu}}^{\dagger }=\tfrac{1}{\sqrt{2}}\left( \widetilde{
\beta }\widetilde{x}^{\widetilde{\mu }}-\widetilde{\beta }^{-1}\partial
/\partial \widetilde{x}^{\widetilde{\mu }}\right)
\end{equation}
and
\begin{equation}
\widetilde{a}_{\tilde{\mu}}=\tfrac{1}{\sqrt{2}}\left( \widetilde{\beta }%
\widetilde{x}^{\widetilde{\mu }}+\widetilde{\beta }^{-1}\partial /\partial
\widetilde{x}^{\widetilde{\mu }}\right) .
\end{equation}
Operating on the $\widetilde{\phi }_{\widetilde{n}_{1}\widetilde{n}_{2}%
\widetilde{n}_{3}\widetilde{n}_{4}}$, the $\tilde{E}_{\tilde{\mu}\tilde{\nu}
} $ satisfy
\begin{equation}
\left[ \tilde{E}_{\tilde{\mu}\tilde{\nu}},\;\tilde{E}_{\tilde{\rho}\tilde{%
\sigma}}\right] =\delta _{\widetilde{\nu }\widetilde{\rho }}\tilde{E}_{%
\tilde{\mu}\tilde{\sigma}}-\delta _{\widetilde{\sigma }\widetilde{\mu }}%
\tilde{E}_{\tilde{\rho}\tilde{\nu}},\;\;\;\tilde{\mu},\;\tilde{\nu},\;\tilde{
\rho},\;\tilde{\sigma}=1,\;2,\;3,\;4.
\end{equation}
But this is the vector-multiplication rule of $SU\left( 4\right) $.
Therefore the fifteen independent $\tilde{E}_{\tilde{\mu}\tilde{\nu}}$
comprise an $SU\left( 4\right) $-algebra and the $\widetilde{\phi }_{%
\widetilde{n}_{1}\widetilde{n}_{2}\widetilde{n}_{3}\widetilde{n}_{4}}$ are
its representations [5, 13]. The smallest representation, with $\widetilde{N}
=0$, is the \textbf{1} consisting of $\;\tilde{\phi}_{0000}=\exp \left( -%
\frac{1}{2}\tilde{\beta}^{2}\tilde{X}^{2}\right) \equiv \tilde{F},$ where $%
\widetilde{X}^{2}=\sum_{\widetilde{\mu }=1}^{4}\left( \widetilde{x}^{%
\widetilde{\mu }}\right) ^{2}$. The next smallest, with $\widetilde{N}=1,$
is the \textbf{4} consisting of\ $\tilde{\phi}_{1000}=\tilde{x}^{1}\tilde{F}
,\;\tilde{\phi}_{0100}=\tilde{x}^{2}\tilde{F},\;\tilde{\phi}_{0010}=\tilde{x}
^{3}\tilde{F},\;$and $\tilde{\phi}_{0001}=\tilde{x}^{4}\tilde{F}.$
Likewise the solutions
\begin{equation}
\widetilde{\phi }_{\widetilde{n}_{1}}\left( \tilde{\beta}\widetilde{x}%
^{1}\right) \widetilde{\phi }_{\widetilde{n}_{2}}\left( \tilde{\beta}%
\widetilde{x}^{2}\right) \widetilde{\phi }_{\widetilde{n}_{3}}\left( \tilde{%
\beta}\widetilde{x}^{3}\right) \equiv \widetilde{\phi }_{\widetilde{n}_{1}%
\widetilde{n}_{2}\widetilde{n}_{3}}
\end{equation}
are basis functions of an $SU\left( 3\right) $-algebra
\begin{equation}
\tilde{E}_{\tilde{i}\tilde{j}}=\widetilde{a}_{\tilde{i}}^{\dagger }%
\widetilde{a}_{\tilde{j}}-\tfrac{1}{3}\delta _{\tilde{i}\tilde{j}}\sum_{%
\widetilde{k}=1}^{3}\widetilde{a}_{\tilde{k}}^{\dagger }\widetilde{a}_{%
\tilde{k}},\;\;\;\tilde{i},\tilde{j}=1,\;2,\;3,
\end{equation}
where $\widetilde{a}_{\tilde{i}}^{\dagger }$ and $\widetilde{a}_{\tilde{j}}$
are defined like $\widetilde{a}_{\tilde{\mu}}^{\dagger }$ and $\widetilde{a}%
_{\tilde{\nu}}$, respectively, except that their indices are restricted to $%
1,\;2,\;$and $3.$ The smallest representation, with $\widetilde{n}_{1}+%
\widetilde{n}_{2}+\widetilde{n}_{3}\equiv \widetilde{N}^{\prime }=0$, is the
\textbf{1} consisting of $\;\tilde{\phi}_{000}=\exp \{-\frac{1}{2}\tilde{%
\beta}^{2}[(\tilde{x}^{1})^{2}+(\tilde{x}^{2})^{2}+(\tilde{x}%
^{3})^{2}]\}\equiv \tilde{F}^{\prime }$. The next smallest, with $\widetilde{%
N}^{\prime }=1,$ is the \textbf{3} consisting of\ $\tilde{\phi}_{100}=\tilde{%
x}^{1}\tilde{F}^{\prime },\;\tilde{\phi}_{010}=\tilde{x}^{2}\tilde{F}%
^{\prime },$ and $\tilde{\phi}_{001}=\tilde{x}^{3}\tilde{F}^{\prime }.$
\end{document}
|
\section{Introduction}
Since the discovery of high-$T_{c}$ (HTC) superconductors their normal state
properties are a subject of extensive investigations. Many of them, as the
electrical resistivity $\rho $, the Hall effect $R_{H}$ or the
thermoelectric power $S$, exhibit some universal behavior in function of
charge carrier doping. The universal behavior of the latter quantity is well
known, for many of the HTC families the thermopower changes sign at optimal
carrier concentration \cite{Obertelli}. The linearity of resistivity for the
optimally doped samples in wide temperature region is also one of their most
well known features \cite{Martin}. The Hall coefficient for optimally doped
samples varies as $1/T$, which results in the $1/T^{2}$ dependence of the
Hall mobility, $\mu _{H}$. Then it appeared that the $1/T^{2}$ dependence of
$\mu _{H}$ is more universal. It applies not only to the optimally doped,
but also to the under- and overdoped materials \cite{Kubo1}.
In this work we present the measurements of the Ettingshausen coefficient,
one of the less known transport coefficients. Our aim was to complement the
present knowledge of the normal state properties of HTC materials, and,
hopefully, to search for some new universalities. The Ettingshausen effect
is a thermal analog of the Hall effect (the definition and sign convention
are shown in Fig. 1). The difference with respect to the Hall effect lies in
the fact that the temperature difference is measured in the direction
perpendicular to both the current and field directions, instead of the
voltage difference.
The Hall and Ettingshausen effects are two of the four transversal
magneto-thermal effects:
\begin{eqnarray*}
\nabla \Phi _{\bot } &=&R_{H}~\vec{j}\times \vec{B}\text{, the Hall effect,}
\\
\nabla T_{\bot } &=&P~\vec{j}\times \vec{B}\text{, the Ettingshausen effect,}
\\
\nabla \Phi _{\bot } &=&Q~\nabla T\times \vec{B}\text{, the Nernst effect,}
\\
\nabla T_{\bot } &=&S_{RL}~\nabla T\times \vec{B}\text{, the Righi-Leduc
effect,}
\end{eqnarray*}
where $R_{H}$, $P$, $Q$, $S_{RL}$ are the respective coefficients, $\nabla
\Phi _{\bot }$ and $\nabla T_{\bot }$ are the transversal gradients of
electrical potential and temperature, respectively, caused by the presence
of magnetic field ($\vec{B}$) perpendicular to the electrical current ($\vec{%
j}$) or heat flux ($\vec{q}\sim \nabla T$). All these coefficients are
interconnected by three fundamental relations which were considered by
P.W.Bridgman \cite{Bridgman} in terms of thermodynamics of reversible
processes:
\begin{eqnarray*}
Q &=&\frac{\kappa }{T}P\text{,} \\
Q &=&\frac{\mu _{T}}{\rho }R_{H}\text{,} \\
P &=&\frac{\mu _{T}T}{\kappa }S_{RL}\text{,}
\end{eqnarray*}
where $\kappa $ is the total thermal conductivity and $\mu _{T}$ is the
Thomson coefficient .
The theory of the Ettingshausen effect for semiconductors was considered in
\cite{Paranjape}, and for metals in \cite{Fieber}. Only few measurements of
the Ettingshausen effect in different materials were carried out up to date.
Typical measured values of the Ettingshausen coefficients for semiconductors
are positive and of the order of 10$^{-2}\div $10$^{-4}$ m$^{3}$K/J (Ge:\cite
{Mette1}; Si:\cite{Mette2}; PbSe, PbTe:\cite{Putley}). Typical values of the
Ettingshausen coefficient for metals are much lower then for semiconductors
and are of the order 10$^{-7}\div $10$^{-8}$ m$^{3}$K/J. A negative
Ettingshausen coefficient was observed for Ag, Cd, Cu, Fe, Zn and Au,
whereas positive for Al, Co and Ni. The Ettingshausen effect and the Hall
effect have opposite signs for Al, Cd, Fe, Ni and Zn, whereas the same signs
for Ag, Co, Cu and Au \cite{CriticalTables,Bridgman}. The Ettingshausen
coefficient is much higher for semi-metals ($\sim $10$^{-4}$ m$^{3}$K/J for
Sb and $\sim $10$^{-3}$ m$^{3}$K/J for Bi \cite{Bridgman}) and for rare
earths ($\sim $10$^{-3}$ m$^{3}$K/J for Y, Gd, Tb and Dy \cite{Zecchina}).
According to our knowledge, no measurements of the Ettingshausen effect were
carried out in HTC superconductors in normal state. Few works were devoted
to the normal state Nernst effect: for Tl-2212 \cite{Clayhold}, YCBO \cite
{Gasumyants} and NdCeCuO \cite{Fournier}.
In the present work we have chosen the La$_{2-x}$Sr$_{x}$CuO$_{4}$ (LSCO)
solid solution ($x=0.03\div 0.35$). This compound exhibits full range of
behaviors versus chemical composition, which is characteristic of the
layered copper-oxide superconductors. The carrier concentration may be
controlled by the Sr concentration $x$. One could therefore investigate the
doping dependence from the semiconducting region ($x\lesssim 0.05$), through
the underdoped ($0.05\lesssim x\lesssim 0.17$) and overdoped ($0.17\lesssim
x\lesssim 0.30$) superconducting regions, up to the heavily doped metallic
region with no superconductivity ($x\gtrsim 0.30$) \cite{Devaux}. Moreover,
LSCO has a simple crystal structure with single CuO$_{2}$ layers. It has
neither Cu-O chains as in YBa$_{2}$Cu$_{3}$O$_{7-\delta }$ nor complicated
modulation of the separating and spacing layers as in Bi- and Tl-based
materials. All our samples have been characterized by X-ray and electrical
resistivity measurements.
\section{Experimental}
\subsection{Samples}
Polycrystalline samples of La$_{2-x}$Sr$_{x}$CuO$_{4}$ were produced
following the standard solid state technique from high purity La$_{2}$O$_{3}$%
, SrCO$_{3}$ and CuO substrates. The powders were mixed and prefired in air
at 950 $^{\circ }$C for 24 h. After pulverization, the materials were mixed,
pressed and sintered at 1000 $^{\circ }$C for 60 h. Then they were
regrounded, pelletized and, except one sample, refired at 1050 $^{\circ }$C
for 72 h in oxygen under pressure of 1 bar. Only the sample of La$_{1.65}$Sr$%
_{0.35}$CuO$_{4}$ was sintered in oxygen under the pressure of 300 bar at
1000 $^{\circ }$C for 48 h. All products were confirmed to be single phase
by powder X-ray diffraction. The values of the critical temperature measured
by the electrical resistivity measurements are shown in Table 1.
Table 1. Critical temperatures of the La$_{2-x}$Sr$_{x}$CuO$_{4}$ samples.
\begin{tabular}{lllllllll}
$x$ (Sr) & 0.03 & 0.05 & 0.10 & 0.15 & 0.20 & 0.25 & 0 30 & 0.35 \\
$T_{c}$ [K] & 0 & 0 & 28.6 & 36.0 & 30.2 & 18.0 & 10.2 & 0
\end{tabular}
\subsection{Choice of the experimental configuration}
The Ettingshausen effect is usually very small, hence it might be easily
overridden by other thermal effects, as the Joule and Thomson effects.
Especially, the Joule heat may particularly hinder the detection of the
Ettingshausen effect. When measuring this effect a strong electrical current
should be passed through the sample thus evolving a large amount of energy
at the electrical contacts, which electrical resistance is usually much
higher than that of the sample. Therefore, some places of the sample should
be thermally anchored to a large mass of constant temperature to carry the
Joule heat away. On the other hand, the sample should be located in
adiabatic conditions since the thermal gradient is the quantity to be
measured. For the above reason we started from computer modelling of
different possible experimental setups, with different patterns of sample
anchoring (e.g. on corners, along the longest side, etc.). In all cases we
have supposed that the sample should have a shape of a flat slab with the
shortest dimension parallel to the magnetic field, since the temperature
difference due to the Ettingshausen effect is inversely proportional to that
dimension (in analogy to the Hall effect). The configuration we have finally
chosen is presented Fig. 2.
The ends of the sample (A) were attached to two copper bars (B) using the
Au:In alloy and silver paint, making both good electrical and thermal
contacts. The typical contact resistance amounted to 0.5$\div $2 $\Omega $
and was much higher than the sample resistance. The bars (B) were
electrically insulated from the copper support (C), but their good thermal
connection to the support was ensured by a relatively large area of the
contacts. The sample was also surrounded by a thick copper screen (D)
screwed on the support (C). The carbon-glass thermometer (E) was \ located
within support (C). The whole assembly was placed in a gas-flow cryostat. A
differential copper-constantan-copper thermocouple was attached to the sides
of the sample. This type of thermocouple has relatively large sensitivity at
room temperature ($\alpha =40.5$ $\mu $V/K). Moreover, since only the middle
segment was made of constantan, the total resistivity of the thermocouple
amounted to few ohms only, reasonable reducing the thermal voltage noise.
Next advantage was that only copper leads were connecting the thermocouple
junctions on the sample with the Keithley 182 nanovoltmeter input, thus
avoiding all detrimental EMF's (the only lead solderings were thermally
anchored to the support C). No detectable influence of the magnetic field on
the thermocouple voltage have been found.
Providing that the length of the sample is much greater than its width, in
our configuration the Ettingshausen coefficient $P$ may be calculated
directly from the definition:
\[
\Delta T_{Ett}=PJ_{x}B_{z}/d
\]
where $\Delta T_{Ett}$ is the temperature difference between both sides of
the sample due to the Ettingshausen effect, $J_{x}$ is the electrical
current, $B_{z}$ is the magnetic induction and $d$ is the sample thickness.
All our samples had thickness of 0.25-0.35 mm, width of 2.5-3 mm and length
of 9-10 mm.
\subsection{Influence of the Joule effect}
The temperature distribution on the sample surface due the Ettingshausen
effect was presented in Fig. 3a. Our calculations have shown that in the
centre of the sample this distribution is independent on the temperature
distribution caused by the Joule effect, which is shown in Fig. 3b. In other
words, the difference $\Delta T_{Ett}$ remains unaffected by the Joule heat,
even if the values of the temperature gradient along the sample are much
higher then the transversal gradient. The reason is that the gradients
produced by the two effects are perpendicular. However, an inevitable
mismatching of the thermocouple junctions positions causes that the Joule
effect also may significantly contribute to the total temperature difference
measured:
\[
\Delta T=T_{2}-T_{1}=\pm \Delta T_{Ett}+\Delta T_{Joule}
\]
Depending on particular mounting, the $\Delta T_{Joule}$ difference may be
of any sign. Its value, which is approximately proportional to $J^{2}$, may
be much higher then $\Delta T_{Ett}$. Fortunately, the $\Delta T_{Ett}$
difference changes its sign upon reversing the direction of both electrical
current and magnetic field, whereas $\Delta T_{Joule}$ difference does not.
This feature was used to extract the Ettingshausen effect from the
background.
\subsection{Influence of the Thomson effect}
There is another effect interfering with determination of the Ettingshausen
temperature difference $\Delta T_{Ett}$ - namely the Thomson effect. Because
of our experimental arrangement a temperature gradient along the sample is
present (see Fig. 3a), so the total heat $q$ produced per time unit in the
sample (without magnetic field) consists of two components:
\[
\dot{q}(x)=J_{x}^{2}\rho -\mu _{T}J_{x}\frac{dT(x)}{dx}
\]
The first term is the Joule heat, the second - the Thomson heat; $\rho $
denotes the electrical resistivity and $\mu _{T}$ - the Thomson coefficient.
According to our knowledge, the values of the Thomson coefficient for HTC
materials are unknown. However, they should be of the order of the Seebeck
coefficient, because of the Kelvin relation, $\mu _{T}=TdS/dT$. The Thomson
effect causes that the temperature distribution along the sample slightly
changes upon changing the direction of the electrical current - see Fig. 4.
Thus, the total temperature difference measured crosswise the sample should
be expressed as:
\[
\Delta T=T_{2}-T_{1}=\pm \Delta T_{Ett}+\Delta T_{Joule}\pm \Delta
T_{Thomson}
\]
The $\Delta T_{Thomson}$ difference changes sign upon changing the direction
of the electrical current, but, fortunately, it is insensitive to the
magnetic field. Hence, the measurements of $\Delta T$ in function of
magnetic field still give a chance to extract the effect of interest, $%
\Delta T_{Ett}$.
In the course of our experiments we have sometimes observed that during
prolonged measurements on the same sample the values of the $\Delta
T_{Joule} $ and $\Delta T_{Thomson}$ changed significantly (including the
sign change for $\Delta T_{Thomson}$), whereas the\ $\Delta T_{Ett}$
remained unaffected. This observation is not surprising since $\Delta
T_{Joule}$ and $\Delta T_{Thomson}$ are just a result of the mismatching of
the thermocouple junctions which can be effectively changed by thermocycling.
\subsection{Influence of the electrical contacts heating and of the gas
cooling}
The resistance of the sample contacts was usually approximately one order of
magnitude higher than that of sample itself, so most of the electrical
energy was released within the contacts. The only consequence for the
temperature distribution within the sample is that the curves shown in Fig.4
are shifted upwards, with no change of both the longitudinal and transversal
temperature gradients. However, due to relatively large electrical current
values we were using (up to 0.5 A) the sample would be heated up by several
kelvins in respect of its surrounding (i.e. the support C with thermometer
and the screen D, see Fig. 2). Therefore, in order to keep the temperature
of the sample possibly close to that of thermometer we decided to perform
all measurements in helium gas atmosphere under ambient pressure, despite
that it will influence all temperature gradients in the sample and its
support. We can assume that the gas has the temperature of the support.
Thus, the gas is colder than the sample heated by the electrical contacts.
Therefore, one could expect that the longitudinal temperature distributions
presented in Figures 3a and 4 would be substantially flatten, or even
inverted, depending on the ratio of the Joule heating to the gas cooling
efficiency. The diminishing of the transversal gradient due to the
Ettingshausen effect depends only on the gas cooling efficiency. Therefore,
it is possible to set the experimental conditions so that the gas cooling is
weak enough to do not disturb the Ettingshausen effect significantly, but to
reduce substantially the detrimental longitudinal gradients. Indeed, in
tests, in which one of the thermocouple junctions was moved onto the bar B,
we have observed that for low values of the measurement current the middle
of the sample may be colder than its ends (see Fig. 5). Moreover, reducing
high temperature gradients by gas cooling should also result in diminishing
the influence of non-linear energy exchange by radiation.
To check the influence of the gas cooling on the results we have performed
several measurements changing pressure of helium gas in the range of 30$\div
$1030 mbar. It has been observed that the value of the measured
Ettingshausen coefficient is insensitive to the pressure within the
measurement error. In contrast, the reduction of the pressure from 1030 mbar
to 30 mbar resulted in increase of the $\Delta T_{Joule}$ difference by
approximately 3 times. The $\Delta T_{Thomson}$ difference was also found to
be quite sensitive to the pressure, even that sign changes were observed.
\subsection{Measurements strategy}
A typical measurements sequence without magnetic field is presented in Fig.
6. For each electrical current value several readings of the thermocouple
voltage were taken, with alternatively changing the current direction. A
delay of 60-120 seconds was applied after each current reversal to allow the
steady-state to build up. The average value of the measured voltage for a
given current value ($\Delta T_{Joule}$) was increasing approximately
proportionally to $J^{2}$, as it might be expected for the Joule effect (see
the upper insert). Changing the current direction resulted in a smaller
voltage variations, which amplitude was found to be dependent on the
electrical current value (see the lower insert). We suppose that the Thomson
effect is responsible for these variations. In this paper we will call this
amplitude $\Delta U_{\pm J}$ (an amplitude of the thermocouple voltage
variations due to reversing of the electrical current). The Thomson
temperature difference may be calculated from the equation:
\[
\Delta U_{\pm J}=\alpha (\Delta T_{Thomson})
\]
where $\alpha $ is the thermocouple sensitivity. As expected from our
computer simulations, we have not observed any universal dependence of $%
\Delta U_{\pm J}$ on electrical current. For some cases $\Delta U_{\pm J}$
changed its sign for a certain value of the current (with no magnetic field).
Application of the magnetic field causes the Ettingshausen effect and now
the amplitude $\Delta U_{\pm J}$ is a sum of two contributions:
\[
\Delta U_{\pm J}=\alpha (\Delta T_{Thomson}+\Delta T_{Ett})
\]
A typical measurement sequence in magnetic field and with constant
electrical current is shown in Fig. 7. The influence of the Ettingshausen
effect is clearly visible. The measured amplitude $\Delta U_{\pm J}$
strongly depends on the field strength. The insert demonstrates the values
of $(\Delta T_{Thomson}+\Delta T_{Ett})=$ $\Delta U_{\pm J}/\alpha $ for
different values of magnetic field fitted by a linear function. The
Ettingshausen coefficient is calculated from the slope $A$ of this line
using the formula:
\[
P=A\frac{d}{J}
\]
The linearity of the dependence of the Ettingshausen effect on electrical
current value may be checked by plotting the differences $\Delta U_{\pm
J}(B)-$ $\Delta U_{\pm J}(B=0)$ versus current (see Fig. 8).
Taking into account all the above considerations we have developed a
measurements strategy which is illustrated in the Fig. 9. For a particular
sample we started from measuring of the current dependencies of the $\Delta
U_{\pm J}$ amplitude at $B$=+8, 0 and -8 T. Hence, we were able to check the
linearity of $\Delta U_{\pm J}$ with the electrical current value. Then, we
chose a value of current for measuring the field dependency of $\Delta
U_{\pm J}$ amplitude. If possible, we chose a value of $J$ for which $\Delta
U_{\pm J}(B=0)$ was close to zero, in order to avoid the influence of
another interfering effect connected with the Righi-Leduc phenomenon, which
is defined as:
\[
\nabla T_{RL}=S_{RL}B_{z}\nabla T_{x}
\]
where $S_{RL}$ is the Righi-Leduc coefficient. This effect may result in
another contribution to the transverse thermal gradient if a longitudinal
thermal gradient is present. Fortunately, the lack of the Thomson effect is
an indication of small longitudinal gradients and in this case the
Righi-Leduc effect may be neglected. Moreover, to reduce additionally the
influence of the Joule and Thomson effects all measurements were performed
at $\sim $1 bar of helium gas.
\section{Results and Discussion}
The method described in this paper allows measurements of Ettingshausen
effect of typical for metals, small value. The effect consists in the
transversal thermal gradient appearing due to the electrical current flow in
the presence of the perpendicular magnetic field. Since this weak thermal
effect may be easily overridden by the Joule and Thomson effects a special
measures have been taken to extract it from the background. The sample ends
were thermally anchored to a large mass to carry away the Joule heat. To
eliminate the influence of the longitudinal thermal gradients due to the
Joule and Thomson effects the odd symmetry of the Ettingshausen temperature
difference in respect with the direction of the magnetic field and
electrical current have been exploited.
The values of the measured Ettingshausen coefficients for La$_{2-x}$Sr$_{x}$%
CuO$_{4}$ are shown in. Fig. 10. For each composition 2 or 3 samples were
investigated. For each sample several measurement runs were carried out and
the results were averaged. All the coefficients are of the order of 10$^{-7}$
m$^{3}$K/J, what is typical value for good metals, in contrast to the
semiconductors, where values higher by several orders of magnitude are
expected (see the Introduction). The Ettingshausen coefficient for La$_{2-x}$%
Sr$_{x}$CuO$_{4}$ changes the sign: it is positive only for low Sr
concentration ($x$ = 0.03 and 0.05), for all other compositions negative
values were observed. This is also in contradistinction to the situation of
semiconductors, where only positive values are predicted by the theory \cite
{Paranjape} and found experimentally (see Introduction). The overall weak
compositional dependence is somehow surprising, since the other transport
coefficients for La$_{2-x}$Sr$_{x}$CuO$_{4}$ were found to be strongly
dependent on Sr/La substitution. It was shown \cite{Suzuki,Takagi} that for
the compositional range $x$ $=0.05\div 0.35$ the Hall coefficient decreases
by more than two orders of magnitude. Similar strong variation with
composition was found for thermoelectric power \cite{Cooper,Goodenough}.
\bigskip
\section{Acknowledgments}
The work was supported by the Polish State Committee for Scientific Research
under contract No. 2PO3B 11613.
\newpage {\LARGE Figure captions}
{\bf Figure1} \ The sign convention for a positive Ettingshausen effect.
{\bf Figure 2} An experimental setup used for measurements of the
Ettingshausen effect. See text for details.
{\bf Figure 3} \ The temperature distribution on the sample area due the
Ettingshausen effect (a) and the temperature distribution along the sample
due to the Joule effect (b). The mismatching of the thermocouple junctions
positions was exaggerated to indicate how the Joule effect may contribute to
the measured temperature difference $\Delta T=T_{2}-T_{1}$. Calculated by
the finite-elements method.
{\bf Figure 4} \ The temperature distribution along the sample due to the
Joule and Thomson effects for two directions of the electrical current. The
difference between two curves was exaggerated for sake of clarity.
Calculated by the finite-elements method.
{\bf Figure 5} \ A realistic temperature distribution along the sample for
different values of the measurement current (calculated by the
finite-elements method). The Joule and Thompson effects within the sample,
the Joule effect in the electrical contacts and the cooling by the
surrounding gas have been taken into account. The sample ends are warmer in
respect with the support due to the Joule heating of the contacts. It is
assumed that the gas has the same temperature as the support. The insert
shows the results of the tests, in which one of the thermocouple junctions
was moved onto the bar B (see Fig. 2) and the longitudinal gradient $\Delta
T_{l/2}$ has been measured.
{\bf Figure 6} \ An exemplary measurement sequence performed on La$_{1.75}$Sr%
$_{0.25}$CuO$_{4}$\ sample at room temperature without magnetic field. The
thermocouple voltages are plotted versus the successive reading numbers. The
direction of the electrical current was alternatively changed from reading
to reading. Our convention is that $J$\ is positive for odd reading numbers,
which are denoted by filled circles. Even numbers are denoted by empty
circles. The upper insert shows the average voltage for each value of the
current versus the current, i.e. the result of the Joule effect (the
thermocouple sensitivity is denoted by $\alpha $). The variations of the
voltage with the current direction are due to the Thomson effect. The
amplitude of these variations $\Delta T_{Thomson}=\Delta U_{\pm J}/2\alpha $
versus current value is shown on the lower insert.
{\bf Figure 7}\ \ An exemplary measurement sequence performed on La$_{1.75}$%
Sr$_{0.25}$CuO$_{4}$\ sample at room temperature for a constant electrical
current and in varied magnetic field (see also the caption of Fig. 6). A
plot of the amplitude ($\Delta T_{Thomson}+\Delta T_{Ett})=\Delta U_{\pm
J}/2\alpha $\ versus $B$\ is shown in the insert.
{\bf Figure 8} \ A plot of the amplitude ($\Delta T_{Thomson}+\Delta
T_{Ett})=\Delta U_{\pm J}/2\alpha $ versus electrical current for different
values of the magnetic field. The insert shows the differences $(\Delta
U_{\pm J}(B)-$\ $\Delta U_{\pm J}(B=0))/2\alpha =\pm \Delta T_{Ett}$ ($%
\alpha $ is the thermocouple sensitivity).
{\bf Figure 9} \ A sketch of the experimental procedure for measurements of
the Ettingshausen effect. At each point on the ($B$,$J$) plane several
readings of the transversal temperature gradient have been taken for both
current directions (see Figures 6 and 7).
{\bf Figure 10} The room temperature Ettingshausen coefficients for La$%
_{2-x} $Sr$_{x}$CuO$_{4}$. The error bars denote the scatter of experimental
values obtained for a particular composition.
|
\subsection*{References:}
\begin{enumerate}
\item Greenstein, J. L. and Trimble, V. L., {\em Astrophysical Journal},
{\bf 149}, 283 (1967); Moffett T. J. et. al., {\em Astronomical Journal},
{\bf 83}, 820 (1978); Shipman, H. L. and Sass, C. A., {\em Astrophysical
Journal}, {\bf 235}, 177 (1980); Grabowski, B. et. al. {\em Astrophysical
Journal}, {\bf 313}, 750 (1987). \label{white}
\item Ghosh, A. in {\em ``Progress in New Cosmologies: Beyond the Big
Bang},'' H. C. Arp et. al. (Ed.), Plenum Press, (1993). \label{AG2}
\item Sinclair, A. T., {\em Astronomy and Astrophysics}, {\bf 220}, 321
(1989). \label{phobos}
\item Sadeh, D., Knowles, S. H. and Yaplee, B. S., {\em Science}, {\bf
159}, 307 (1968). \label{sadeh}
\item Merat, P., Pecker, J-C. and Vigier, J-P., {\em Astronomy and
Astrophysics}, {\bf 174}, 168
(1974). \label{merat}
\item Ghosh, A., {\em Pramana (Journal of Physics)}, {\bf 26}, 1 (1986).
\label{AG1}
\end{enumerate}
\end{document}
|
\section{Introduction}
The dynamics of a system undergoing phase-ordering following a quench
from the high temperature (disordered) phase to the ordered phase is of great
interest\cite{Review}. The kinetics of systems with $O(n)$ symmetry
subject to `Model A' dynamics \cite{HH}(i.e.\ systems with nonconserved order
parameter) and `Model B' dynamics \cite{HH} (systems with conserved order
parameter) have been previously studied \cite{BH1,FA} within a Gaussian closure
theory originally developed by Mazenko \cite{MAZ1,MAZ2} following the
seminal work of Ohta, Jasnow and Kawasaki (OJK) \cite{OJK}. In previous work
\cite{BRC} we have computed the form of the corrections to the scaling limit,
and the correction-to-scaling exponent, for a number of systems with
nonconserved order parameter. These include some exactly soluble models, and
the Model A dynamics of a scalar field within the Mazenko theory.
In the present work we turn our attention to systems with a vector order
parameter, both nonconserved and conserved. The corrections to scaling for
systems with continuous symmetry will be calculated using the Mazenko theory.
It should be mentioned that this approach has been shown to be more successful,
at a quantitative level, in systems with nonconserved order parameter than
those with conserved order parameter \cite{RPB}. Nevertheless, the results
obtained in the conserved case are in qualitative agreement with those obtained
in simulations. Furthermore, the Mazenko approach seems the only available
method to probe the questions of corrections to scaling addressed here.
In particular, we found that, for nonconserved scalar fields, the
correction-to-scaling exponent $\omega$ is predicted by this approach to have
a nontrivial value. We will show that this same feature is present for the
vector fields, with and without conservation.
It is well established \cite{Review} that at late times most phase-ordering
systems approach a scaling regime, where the equal-time pair correlation
function $C(r,t) \equiv \langle\vec{\phi}({\bf x}+{\bf r},t)\cdot
\vec{\phi}({\bf x},t)\rangle$ takes the form $C(r,t)= f[r/L(t)]$.
The characteristic length scale $L(t)$ grows with time as
$L(t) \sim t^a$, where $a$ is the growth exponent which depends on the
nature of the dynamics and the symmetry of the order parameter.
In particular, $a = 1/2$ for nonconserved order parameter systems, while
$a = 1/3$ for systems with conserved scalar order parameter and $a = 1/4$
for systems with conserved vector order parameter (with logarithmic
corrections for $n = 2$, $d > 2$ \cite{BR}). In previous work \cite{BRC}
we studied how scaling is approached in nonconserved order
parameter models such as the 1-$d$ Ising model with Glauber dynamics, the
$n-$vector model with $n = \infty$, the approximate OJK theory and the
Mazenko theory for scalar fields. In all these cases $\omega$ was found to
be trivial ($\omega = 4$) except the last, for which $\omega$ was found to be
non-trivial and dimensionality dependent.
The relevance of corrections to scaling lies in interpreting
experimental and simulation results, where it is advantageous to know how
the scaling limit is approached. Corrections to scaling in systems with finite
$n>1$ in $d =3$ and $d = 2$ were not considered in \cite{BRC}.
The main objective in this article is to study systems with $n \ge 2$.
This article is devoted to the study of the corrections to scaling
for systems with $O(n)$ symmetry in phase-ordering dynamics. The leading
corrections to scaling enter the correlation function in the form
\begin{equation}
C(r,t) = f_0(r/L)+ L^{-\omega}f_1(r/L),
\end{equation}
where $f_0(x)$ is the `scaling function' and $f_1(x)$ the
`correction-to-scaling function'. The quantity which unites theory, computer
simulation and experiment is the structure factor
$S(k,t) = L^d g_{0}(y) + L^{d - \omega} g_{1}(y)$,
where $g_{0}(y)$ and $g_{1}(y)$ are the $d$-dimensional Fourier transforms of
$f_{0}(x)$ and $f_{1}(x)$ respectively, and $y=kL$. Coniglio and Zannetti
\cite{CZ} solved the conserved $O(n)$ model for $n = \infty$ exactly, and
found that no simple scaling exists. Instead a `multiscaling' behavior was
obtained, raising the question of whether simple scaling exists in conserved
order parameter systems with $n > 1$ generally (or even for conserved scalar
fields). However, it was later shown by Bray and Humayun \cite{BH2},
analytically within the Mazenko theory, that scaling does exist for large
but finite $n$. Attempts to find multiscaling behaviour in simulation date
for conserved scalar fields \cite{CZ2}, or the conserved XY model in two
\cite{MG} or three \cite{SR} dimensions were not successful. It is now
generally believed that scaling is recovered asymptotically in time in the conserved
$O(n)$ model, for all finite $n$, though multiscaling may be observable in
the preasymptotic regime \cite{Zannetti}.
There are other sources of corrections to scaling apart from the one
considered in this paper. In phase-ordering systems there is, in addition
to the time-dependent coarsening scale $L(t)$, a second
characteristic length scale -- the `defect core size' $\xi$ -- in systems
with topological defects. The corrections to scaling associated with nonzero
defect core size (where $\xi$ is the domain wall thickness in scalar systems)
are expected to enter as a power of ${\xi}/L$. Here we are interested
primarily in the corrections to scaling associated with non-scaling initial
conditions. We therefore suppress the contributions associated with nonzero
core size $\xi$ by taking the `hard-spin' limit, i.e.\ working with an
order-parameter field whose length is everywhere unity, $\vec{\phi}^2=1$,
which forces $\xi=0$ (though in the Mazenko theory this limit will be taken
at the end). Also thermal fluctuations at $T > 0$ may give rise to
significant corrections to scaling for systems quenched to a nonzero final
temperature $T$ \cite{Bray89} (where $0 < T < T_{c}$, with $T_{c}$ the critical
temperature) as has been shown explicitly in the nonconserved $O(n)$ model with
$n \to \infty$ \cite{BN}. However, we will only be studying systems quenched
to $T = 0$. Although corrections to scaling due to thermal fluctuations and
nonzero $\xi$ are important we will not consider them further in this paper.
The outline of the paper is as follows. In the following section
the approximate Mazenko theory is discussed and some general concepts are
introduced. Section III deals with nonconserved order parameter systems.
In section IV, corrections to scaling for the nonconserved 1-$d$ XY model
will be studied. Systems with conserved order parameter are considered in
section V. Section VI concludes with a summary and discussion.
\section{Mazenko Theory}
A `Gaussian closure' theory, building on the earlier work of Ohta, Jasnow
and Kawasaki \cite{OJK} has been developed by Mazenko \cite{MAZ1}. This theory
has been successfully applied to $O(n)$ models in the theory of phase-ordering
dynamics \cite{BH1,RPB}. The equation of motion for an order parameter $\vec\phi$
with continuous symmetry, for systems quenched to $T = 0$, is
\begin{equation}
\frac {\partial {{\vec \phi}(1)}}{\partial t_{1}} =
(-\nabla^2_1)^p \left[\nabla^2_1{\vec \phi}(1)-
\frac{\partial V[{\vec \phi}(1)]}{\partial {\vec \phi}(1)} \right],
\label{OPE}
\end{equation}
where $p=1$ and $p=0$ for conserved order parameter (Model B) and
nonconserved order parameter (Model A) systems respectively. In (\ref{OPE}),
$V({\vec \phi})$ is a symmetric double-well potential for the scalar case,
and a `wine bottle' potential with a degenerate continuum manifold for a vector
order parameter. Compact notation has been used in which
`1' represents the space-time point $({\vec x}_1,t_1)$ and $\nabla^2_1$
means the Laplacian with respect to ${\vec x}_1$. Multiplying (\ref{OPE})
by ${\vec \phi}(2)$, averaging over initial conditions, and using the
translational invariance of $C(12)$ gives (for $t_1 = t_2 =t)$
\begin{equation}
\frac{1}{2}\,\frac {\partial {C(12)}}{\partial t} =
(-{\nabla^2})^p \left[\nabla^2 C(12)-
\langle \frac{\partial V[\vec\phi(1)]}{\partial\vec\phi(1)}
\cdot \vec\phi(2) \rangle \right],
\label{CFE}
\end{equation}
where now $\nabla^2$ is the Laplacian with respect to
$r = |{\bf x}_{1} -{\bf x}_{2}|$ and $C(12) =
\langle {\vec \phi}(1)\cdot{\vec \phi}(2) \rangle $.
The angular brackets denote the average over the initial conditions.
In order to evaluate the average of the last term in (\ref{CFE}) one
introduces an auxiliary field ${\vec m}(r,t)$ related to
$\vec \phi$ by $\nabla^2_m \vec \phi =
2\,\partial V(\vec \phi)/\partial\vec\phi$, with boundary
condition $ {\vec \phi} \to \vec{m}/|\vec{m}|$ as $|\vec m| \rightarrow
\infty$, and $\vec\phi=0$ at $\vec{m}=0$. Near a defect, the field
$\vec{m}({\bf r})$ is the position vector of the point ${\bf r}$ in
the plane normal to the defect. The assumption that ${\vec m}$ is a
Gaussian field enables the evaluation of the average of the last term
on the right hand side of (\ref{CFE}) giving \cite{BH1}
\begin{equation}
\frac{1}{2}\,\frac {\partial {C(12)}}{\partial t} =
(-\nabla^2)^p \left[\nabla^2 C(12)+
\frac{1}{2S_{0}(1)}\,\gamma\,\frac{dC(12)}{d\gamma}\right],
\label{ACFE}
\end{equation}
where $S_{0} = \langle m(1)^2 \rangle$ and ${\gamma}(12) =
\langle m(1)m(2) \rangle /[\langle m(1)^2 \rangle \langle m(2)^2
\rangle]^{1/2}$ is the normalised correlator of the field $m$ (where $m$
is one of the components of ${\vec m}$). An explicit
expression which relates $\gamma$ to $C(12)$ was given in \cite{BPTO}
\begin{equation}
C = \frac{n{\gamma}}{2{\pi}}\,
\left[B\left(\frac{n+1}{2},\frac{1}{2}\right)\right]^2\,
F\left(\frac{1}{2},\frac{1}{2};\frac{n+2}{2};{\gamma}^2\right),
\label{BPT}
\end{equation}
where $B(y,z) = \Gamma(y)\Gamma(z)/\Gamma(y+z)$ is the Beta function
and $F(a,b;c;z)$ the hypergeometric function. Equations (\ref{ACFE}) and
(\ref{BPT}) provide closed form equations for $C(12)$. On substituting
(\ref{BPT}) in (\ref{ACFE}) one obtains an equation for $\gamma$ which can
in principle be solved numerically and substituted back into (\ref{BPT})
to obtain the correlation function $C(12)$. We note at this point that
in deriving the correlation function (\ref{BPT}), the `hard-spin' limit
$\phi = \vec{m}/|\vec{m}|$ was employed. Since this result holds
far from defect cores, it will correctly describe the scaling limit where
the defects are dilute. Here we are also using it to compute the corrections
to scaling.
\section{Nonconserved O(n) Model}
For a nonconserved system $p = 0$, and equation (\ref{ACFE}) is simply
\begin{equation}
\frac{1}{2}\,\frac {\partial {C(12)}}{\partial t} =
{\nabla^2}C(12)+ \frac{1}{2S_{0}(1)}\,\gamma\,\frac{dC(12)}{d\gamma}
\label{CCCE}
\end{equation}
For $n = 1$, using the properties of the hypergeometric function the last
term on the right hand side of (\ref{CCCE}) can be written in terms of
$C(12)$ only, resulting in an equation which is independent of ${\gamma}(12)$.
Corrections to scaling in this case where obtained in our previous work
\cite{BRC}, and will not be considered further here. For general $n$,
$\gamma$ cannot be eliminated in favour of $C(12)$, and we will
therefore work with $\gamma$ instead of $C(12)$. From dimensional
considerations we see that $S_{0} \sim L^2$ and can be chosen as $S_{0} =
{L^2}/{\lambda}$. This choice effectively defines $L$, up to an overall
constant. For $n \rightarrow \infty$, an expansion in $1/n$ can be
performed on $C(\gamma)$, and in this limit $\gamma\,dC/d\gamma =
C + {C^3}/n + O(1/n^2)$. For $n = \infty$, Mazenko theory reduces to the
$n = \infty$ $n$-vector model for which an exact solution, including the
corrections to scaling, is known \cite{BRC}. Expressing (\ref{CCCE}) in
terms of ${\gamma}$ explicitly leads to
\begin{equation}
\frac{1}{2}\,\frac {\partial {\gamma}}{\partial t} =
\frac{C_{\gamma \gamma}}{C_{\gamma}}\,\left(\frac{\partial \gamma}
{\partial r}\right)^2 + \frac{\partial^2 \gamma}{\partial r^2}
+\frac{d-1}{r}\,\frac {\partial {\gamma}}{\partial r}+
\frac{\lambda}{2L^2}\,\gamma\ ,
\label{CCE}
\end{equation}
where $C_{\gamma} = dC/{d{\gamma}}$ etc. Since $C(r,t)$ is a function of
$\gamma(r,t)$, the scaling and corrections to scaling can be imposed on
$\gamma(r,t)$. In the scaling limit we expect $\gamma(r,t)$ to approach
the scaling function ${\gamma_{0}}(r/L)$ which is $L$-independent if all
lengths are scaled by $L$. In this limit therefore one expects
$LdL/dt = $ constant. Including corrections to scaling in $\gamma(r,t)$
and $L(t)$ as usual \cite{BRC} we can write
\begin{eqnarray}
\gamma(r,t) & = & \gamma_0\left(\frac{r}{L}\right) + L^{-\omega}\,
\gamma_1\left(\frac{r}{L}\right) + \cdots, \\
C(r,t) & = & f_0\left(\frac{r}{L}\right)
+ L^{-\omega}\,f_1\left(\frac{r}{L}\right) + \cdots, \\
\frac{dL}{dt} & = & \frac{1}{2L} + \frac{b}{L^{1+\omega}} + \cdots,
\label{L}
\end{eqnarray}
where
\begin{eqnarray}
\label{S}
f_0\left(\frac{r}{L}\right) &=& C(\gamma_0), \\
f_1\left(\frac{r}{L}\right) &=& {\gamma_{1}}\left(\frac{r}{L}\right)\,
{\left[\frac{dC}{d\gamma}\right]}_{\gamma = \gamma_0},
\label{CS}
\end{eqnarray}
and $b$ is a constant. Equating leading and next-to-leading powers of $L$
in the usual way gives
\begin{equation}
{\gamma_0}'' + \frac{C_{\gamma_{0} \gamma_{0}}}{C_{\gamma_{0}}}\,
{{\gamma_{0}}'}^2 + \left[\frac{x}{4}+\frac{d-1}{x} \right ]\,{{\gamma_{0}}'}
+\frac{\lambda}{2}\,\gamma_{0}=0,
\label{NSE}
\end{equation}
\begin{eqnarray}
{\gamma_1}'' + \left[\frac{x}{4}+\frac{d-1}{x} \right ]\,
{{\gamma_{1}}'}+\left [\frac{\lambda}{2}+\frac{\omega}{4} \right]\,
\gamma_{1} + \frac{b}{2}\,x\,{{\gamma_{0}}'} + \nonumber\\
2\,\frac{C_{\gamma_{0} \gamma_{0}}}{C_{\gamma_{0}}}\,
{{\gamma_{0}}'}\,{{\gamma_1}'}
+ \left [ \frac{C_{\gamma_{0} \gamma_{0} \gamma_{0}}}{C_{\gamma_{0}}}
- \frac{(C_{\gamma_{0} \gamma_{0}})^2}
{{C_{\gamma_0}}^2} \right]\,\gamma_{1}\,{{\gamma_0}'}^2=0,
\label{NCSE}
\end{eqnarray}
with $C_{\gamma_{0}} = [dC/d\gamma]_{\gamma=\gamma_0}$ etc. The primes
indicate derivatives with respect to the scaling variable $x = r/L$.
Equations (\ref{NSE}) and (\ref{NCSE}) are to be integrated numerically
subject to appropriate `initial' conditions imposed at $x=0$. Since $x=0$
corresponds to $\gamma_0=1$, the initial conditions are obtained by
considering the regime ${\gamma_0} \rightarrow 1$. Using the properties
of the hypergeometric functions\cite{ABS} one can derive relations between
$C({\gamma_0})$ and its derivatives as ${\gamma_0} \rightarrow 1$.
Up to prefactors of order unity, we find in this limit
\begin{eqnarray}
C_{\gamma_0\gamma_0}/C_{\gamma_0} & \sim & [(1-\gamma_0)
|\ln (1-\gamma_0)|]^{-1},\ \ \ n=2 \nonumber \\
C_{\gamma_0\gamma_0\gamma_0}/C_{\gamma_0} & \sim &
[(1-\gamma_0)^2|\ln(1-\gamma_0)|]^{-1}\, \ \ \ n = 2 \nonumber\\
C_{\gamma_0\gamma_0}/C_{\gamma_0} & \sim & [1-\gamma_0]^{(n-4)/2},
\ \ \ 2<n<4 \nonumber \\
C_{\gamma_0\gamma_0\gamma_0}/C_{\gamma_0} &\sim &
[1-\gamma_0]^{(n-6)/2},\ \ \ 2 < n < 4 \nonumber\\
C_{\gamma_0\gamma_0}/C_{\gamma_0} &\sim&
|\ln(1-\gamma_0)|, \ \ \ n=4 \nonumber\\
C_{\gamma_0\gamma_0\gamma_0}/C_{\gamma_0}&\sim & [1-\gamma_0]^{-1},
\ \ \ n = 4 \nonumber\\
C_{\gamma_0\gamma_0}/C_{\gamma_0} & \to & {\rm constant},
\ \ \ 4 < n < 6 \nonumber\\
C_{\gamma_0\gamma_0\gamma_0}/C_{\gamma_0}&\sim &
[1-\gamma_0]^{(n-6)/2},\ \ \ 4 < n < 6,
\label{SMLC}
\end{eqnarray}
and so on. We have given explicit expressions for $C_{\gamma_0\gamma_0}
/C_{\gamma_0}$ and $C_{\gamma_0\gamma_0\gamma_0} /C_{\gamma_0}$ as
$\gamma_0 \rightarrow 1$ for the values of $n$ which we are going to study.
Using the above results one can show \cite{BH1} that the small-$x$ behavior
of $\gamma_0(x)$ is given by
\begin{equation}
\gamma_{0}(x) = 1- \frac{\lambda}{4d}\,x^2 + \cdots
\end{equation}
for $n \ge 2$, where the limiting forms in (\ref{SMLC}) were used to
demonstrate that the term involving $C_{\gamma_0\gamma_0}/C_{\gamma_0}$
in (\ref{NSE}) is subdominant as $x \to 0$ for $n \ge 2$.
For large-$x$, $\gamma_{0} \rightarrow 0$ (also $C(12) \rightarrow 0$) and
equation (\ref{NSE}) becomes linear because in this limit the second term
in (\ref{NSE}) is negligible. It is easy to show that two linearly
independent solutions of the linearised equation have the asymptotic
forms $\gamma_{01} \sim x^{-2\lambda}$ and
$\gamma_{02} \sim x^{2\lambda - d}\exp(-x^2/8)$, for $x \to \infty$.
As equation (\ref{NSE}) is integrated forward from $x = 0$, the large-$x$
solution obtained will in general be a linear combination of $\gamma_{01}$
and $\gamma_{02}$. The amplitudes of $\gamma_{01}$ and $\gamma_{02}$, however,
depend on $\lambda$. For systems with initial conditions containing only
short-range spatial correlations (as is the case for systems quenched from
high temperature), a power-law decay is unphysical, and $\lambda$ is
determined by the condition that the coefficient of the power-law term,
$\gamma_{01}$, must vanish \cite{BH1}. Note that $\lambda$ is related to the
exponent $\bar{\lambda}$ describing the decay of the autocorrelation function
\cite{LIMAZ} via ${\bar\lambda} = d - \lambda$. Values for $\lambda$ are
given in Table 1 for $2 \le n \le 5$ and $1 \le d \le 3$. Comparison
of the predicted values of $\lambda$ with simulations \cite{NBMH} and
experiments \cite{MPY} show reasonable agreement. It can be shown
that for $d \rightarrow \infty$ the OJK result ($\lambda = d/2$) is
recovered for both scalar \cite{LiuMaz} and vector \cite{BH1} cases.
The same limit for $\lambda$ is also obtained for $n \rightarrow \infty$
at arbitrary $d$.
The correction-to-scaling exponent, $\omega$, is found from (\ref{NCSE})
in a similar way to the determination of $\lambda$ from (\ref{NSE}).
In order to specify initial conditions for the numerical integration of
(\ref{NCSE}), we need the small-$x$ behavior of $\gamma_1(x)$. A small-$x$
analysis of (\ref{NCSE}) gives $\gamma_1 = b\lambda x^4/16d(d+2) + \cdots$,
where the results in (\ref{SMLC}) were used to show that the last two
terms in (\ref{NCSE}) are subdominant as $x \to 0$. The required initial
conditions are therefore $\gamma_1(0) = \gamma_1'(0) = 0$.
As $x \rightarrow \infty$, the last two terms in (\ref{NCSE}) can be
neglected. The two linearly independent solutions of the simplified
equations have a power-law tail ($\sim x^{-(\omega + 2{\lambda})}$) and
a Gaussian tail ($\sim x^q\exp(-{x^2}/8$)) for large $x$, where
$q=2\lambda-d+\omega$ if $\omega>2$ and $q=2\lambda-d+2$ otherwise. Having
already found $\lambda$, $\omega$ is chosen on physical grounds in the
same way as $\lambda$, namely that the coefficient of the power-law term
in the large-$x$ solution should vanish. The values of $\omega$ obtained
are given in Table 2 for $2 \le n \le 5$ and $1 \le d \le 3$. Note that
$\omega \to 4$ for $d \rightarrow \infty$, as the OJK result (and its
generalization to vector fields) is recovered in this limit.
After solving (\ref{NSE}) and (\ref{NCSE}) for $\gamma_{0}(x)$
and $\gamma_{1}(x)$ we use these results to get the scaling function
$f_{0}(x)$ and the correction-to-scaling function $f_{1}(x)$ from
Eqs.\ (\ref{S}) and (\ref{CS}). Figure 1
shows the scaling functions $f_{0}(x)$ and the correction-to-scaling
functions $f_{1}(x)$ for $n = 2$ and 3 in 3$d$. The amplitude of
$f_{1}(x)$ is arbitrary. It is determined by the coefficient $b$
introduced in (\ref{L}): the value $b = 2$ was used in Figure 1. The
scaling functions and the correction-to-scaling functions do not show strong
dependence on $n$ and $d$ for $n \ge 2$. For $n = 1$ and $n \ge 2$ the
scaling functions are very different, especially in the small$-x$ region. The
reason for this is the presence of the sharp interfaces in $n = 1$ systems,
which lead to a finite slope at the origin in $f_0(x)$ \cite{Review} and
a cubic small-$x$ behaviour in $f_1(x)$ \cite{BRC}. For $n \ge 2$, the
small-$x$ is quadratic for $f_0$, and quartic for $f_1$, with logarithmic
corrections for even $n$.
Within Mazenko theory the correction-to-scaling exponent $\omega$ is
non-trivial and depends on both $n$ and $d$, with $\omega \le 4$ for
all $n$ and $d$ in nonconserved $O(n)$ models. The upper bound of $4$
is obtained when $d \rightarrow \infty$ (for any value of $n$) or
$n \rightarrow \infty$ (for any value of $d$).
A noteworthy feature of Figure 1 is that the correction-to-scaling function
$f_1(x)$ is much larger than $f_0(x)$ at large $x$ (the same feature was
found for many of the models studied in \cite{BRC}). This means that, in
fitting data, scaling violations at large-$x$ should be given less weight
in choosing fitting parameters (e.g.\ the scale length $L(t)$) than
violations at small or intermediate $x$, because corrections to scaling are
larger there.
\section{The one-dimensional XY model}
An exact solution of this model was first presented by Newman et al\cite{NBMH}.
The solution yields an `anomalous' growth law, $L \sim t^{1/4}$, for the
characteristic length exhibited by the pair correlation function,
compared with the usual $L \sim t^{1/2}$ growth law of nonconserved
models. Mazenko theory does not predict this growth law, for the simple
reason that the theory has been built in such way that it might be expected
to give qualitatively correct results only for systems with topological
defects (i.e. $n \le d$), since the $n$-component auxiliary field
${\vec m}({\bf r},t)$ is defined in terms of the underlying defect
structure. Despite this, the theory does a reasonable job of accounting
for the behavior of systems with $n>d$, and in fact becomes exact in the
limit $n \to \infty$. However, systems with $n=d+1$, which can support
topological textures, are poorly treated by this approach.
An exact solution for the nonconserved $n=2$, $d=1$ system is possible
because the equation of motion for the order parameter becomes linear in
the angle representation, $\vec\phi = (\cos\theta,\sin\theta)$, which is
natural in the hard-spin limit, where $\vec\phi^2 = 1$. In this limit
the free energy functional is simply
$F = (1/2) \int dx (d\vec\phi/dx)^2 = (1/2)\int dx (d\theta/dx)^2 $.
The zero-temperature equation of motion for model A,
$\partial\vec\phi/\partial t = -\delta F/\delta\vec\phi$, becomes,
in the angle representation,
\begin{equation}
\frac{\partial\theta}{\partial t} = \frac{\partial^2\theta}{\partial x^2},
\label{NDE}
\end{equation}
which is a diffusion equation for the phase angle $\theta$. Thus one
characteristic length scale is the `phase diffusion length',
$L_\theta = t^{1/2}$, but this is not the scale which characterises the
pair correlation function.
Equation (\ref{NDE}) can be solved in Fourier space to give
${\theta_{k}}(t) = {\theta_{k}}(0)\exp(-{k^2}t)$. In evaluating quantities
of interest such as correlation functions, one needs to specify the initial
conditions. The probability distribution, $P([\theta_k(0)])$, for
$\theta_k(0)$ is conveniently chosen to be Gaussian
$P([\theta_k(0)]) \propto \exp
\left[-\frac{1}{2}\sum_k\beta_k\theta_k(0)\theta_{-k}(0) \right ]$.
The choice $ \beta_k = \xi_0k^2/2$ is made as it gives the initial condition
$C(r,0) = \exp(-r/{\xi_{0}})$ for the order-parameter correlation function,
which is the appropriate form for systems quenched from an equilibrium
disordered state with correlation length $\xi_{0}$. The equal-time
correlation function is given by
\begin{equation}
C(r,t) = \langle {\vec \phi}(x,t) \cdot {\vec \phi}(x+r,t)\rangle
=\langle \cos[\theta(x+r,t) - \theta(x,t)]\rangle.
\label{NMCFE}
\end{equation}
Using the Gaussian probability distribution for $\theta_{k}(0)$ equation
(\ref{NMCFE}), with the dynamics (\ref{NDE}) gives
\begin{equation}
C(r,t) = \exp \left(-\sum_{k} \frac{1}{\beta_{k}}\exp(-2{k^2}t)[1-\cos{kr}]
\right).
\label{NMCWE}
\end{equation}
Since the characteristic value of $k$ in the integral is of order
$t^{-1/2}$, and we anticipate (see below) the growth law $L(t) \sim t^{1/4}$,
which sets the characteristic scale of $r$ in $C(r,t)$, it follows
that the scaling limit and the corrections to it can be obtained
from a power-series expansion of $\cos(kr)$ in (\ref{NMCWE}), since the
characteristic value of $kr$ is small (of order $t^{-1/4}$) at late times.
Retaining the leading and next-to-leading terms in the exponent, and
evaluating the sums over $k$, gives
\begin{eqnarray}
C(r,t) & = & \exp \left[-\frac{r^2}{2\xi_0 (2\pi t)^{1/2}} +
\frac{r^4}{96\xi_0 (2\pi)^{1/2}t^{3/2}}
+ O\left(\frac{r^6}{\xi_0 t^{5/2}} \right) \right] \nonumber \\
& = &\exp \left[-\frac{y^2}{2(2\pi)^{1/2}} \right]\,\left[1 +
\frac{\xi_0^2}{L^2}\,\frac{y^4}{96(2\pi)^{1/2}} +
O\left(\frac{\xi_0^4}{L^4}\right) \right],
\label{NMCE}
\end{eqnarray}
where $y = r/L$ is the scaling variable and the coarsening scale
$L(t) = \xi_0^{1/2}t^{1/4}$.
The correction-to-scaling exponent is $\omega = 2$.
This growth law is rather unusual since the generic form of the growth law
for nonconserved fields is $L(t) \sim t^{1/2}$. In this model $\omega$ is
found to be trivial while within Mazenko theory $\omega$ is non-trivial.
There are two fundamental length scales in this problem, namely the phase
coherence $t^{1/2}$ and the correlation length $\xi_0$ associated with the
initial conditions. The coarsening scale $L(t)$ of the pair correlation
function is the geometric mean of these two lengths.
Note that the pair correlation function has a strong dependence on $\xi_0$,
which is not `forgotten' at late times. This sensitivity to initial
conditions is absent in other models, such as
$ n = \infty$ vector model, where the initial conditions drop out at late
times. In the conserved 1-$d$ XY model also, simulation
results gives $L \approx t^{1/6}$ \cite{MG} instead of the
$L \sim t^{1/4}$ behavior expected in higher dimensions \cite{BR},
suggesting that this `anomalous' behaviour may be present
there also. Although no exact solution is known for the conserved case,
heuristic arguments, based on the role of the two characteristic lengths,
can account for the observed $t^{1/6}$ growth \cite{RB}.
\section{Conserved O(n) model}
The dynamical scaling properties of systems with a conserved order
parameter (Model B) with $O(n)$ symmetry is studied using Mazenko theory.
Naive application of this theory does not give correct growth
law $L \sim t^{1/3}$ for scalar fields (the bulk diffusion field must be
included in order to get the correct law\cite{Maz}). Here, however, we
will only consider systems with $n \ge 2$. For Model B systems, equation
(\ref{ACFE}) becomes
\begin{equation}
\frac{1}{2}\,\frac {\partial C(12)}{\partial t} =
-\nabla^2\left[\nabla^2 C(12)+
\alpha (t)\,\gamma\,\frac{dC(12)}{d\gamma}\right]\ ,
\label{CCFE}
\end{equation}
with $\alpha(t) = 1/2S_0$. For eq. (\ref{CCFE}) to have a scaling solution
it is clear that $\alpha\sim 1/L^2$ and $L \sim t^{1/4}$. The latter
is the correct growth law for $n \ge 2$, but for $n = 2$, $d > 2$ there are
logarithmic corrections\cite{BR} which (\ref{CCFE}) fails to predict.
We will first consider the case where $n$ is very large. In this case an
expansion in $1/n$ can be made in equation (\ref{CCFE}). For large $n$,
$C(\gamma)\sim\gamma-\gamma(1-\gamma^2)/2n + O(1/n^2)$ and $\gamma\,
\frac{dC(12)}{d\gamma} = C + C^3/n + O(1/n^2)$. With the above truncations,
(\ref{CCFE}) can be written as
\begin{equation}
\frac{1}{2}\,\frac {\partial C(12)}{\partial t} =
-\nabla^2\left[\nabla^2 C(12)+
\alpha(t)\,\left(C + \frac{C^3}{n} \right) \right]\,,
\label{LNSE}
\end{equation}
correct to order $1/n$.
It is worth mentioning that the $C^3/n$ term is essential for scaling
to be recovered at finite $n$. For $n$ strictly infinite the $C^3/n$ term
is absent and `multiscaling' is obtained \cite{CZ}. For arbitrary $n$, an
expansion in powers of $C$ can be made. Truncating the expansion at order
$C^3$ leads back to (\ref{LNSE}) but with $n$ replaced by an effective
$n^*$, given by $n^* = (n+2)\/a_n^2$ with
$a_n = n[B((n+1)/2,1/2)]^2/2\,\pi$ \cite{FA}.
Dimensional analysis of (\ref{LNSE}) requires $\alpha(t) = \alpha/L^2$,
which defines $L$. Including the leading corrections to scaling as usual
we write
\begin{eqnarray}
C(r,t) & = & f_0(r/L) + L^{-\omega} f_1(r/L) + \cdots \\
dL/dt & = & 1/4{L^3} + b/L^{\omega + 3} + \cdots\ ,
\end{eqnarray}
where $b$ fixes the amplitude of $f_1(r/L)$. Inserting these expansions
into (\ref{LNSE}) and comparing terms of leading order, $O(1/L^4)$, and
next-to-leading order, $O(1/L^{(4 + \omega)})$, gives
\begin{eqnarray}
\label{LNSF}
\frac{x}{8}\,\frac {df_0}{dx} & = &
\nabla_x^2 \left[\nabla_x^2 f_0 +
\alpha\,\left(f_0 + \frac{f_0^3}{n} \right) \right] \\
\frac{x}{8}\,\frac {df_1}{dx} + \frac{\omega}{8}\,f_1
+ \frac{bx}{2}\,\frac{df_0}{dx} & = &
\nabla_x^2\left[\nabla_x^2 f_1 +
\alpha\,\left(f_1 + \frac{3 f_0^2 f_1}{n} \right) \right]\ ,
\label{LNCF}
\end{eqnarray}
where $\nabla_x^2 = d^2/dx^2 + [(d - 1)/x]\,d/dx$.
For general $n$ one must solve (\ref{CCFE}) with $C(r,t)$ given by
(\ref{BPT}). However, the singularities of $C(\gamma)$ and its
derivatives at $\gamma = 1$ introduce some numerical difficulties.
Instead, therefore, we solve (\ref{LNSE}) which is valid for large $n$.
For general $n$, an expansion in $C$ up to $C^3$, leading to (\ref{LNSE})
with an effective $n^*$ \cite{FA}, gives scaling functions
which are in fairly good agreement with simulation results \cite{FA,SR}.
In solving (\ref{LNSF}) numerically, one must know the boundary
conditions. These are provided by small-$x$ and large-$x$ analyses.
For small-$x$, the series expansion $f_0 = 1 +
\sum_{r=1}^\infty \beta_r x^r$, substituted into (\ref{LNSF}),
gives $f_0 = 1 + \beta x^2 - (1+3/n)\,\alpha\,\beta\,
x^4/{4(2+d)} + \cdots$, with $\beta_{2} = \beta$. Numerical integration
can therefore be performed on (\ref{LNSF}) with initial conditions $f_0(0)
= 1$, $f_0''(0) = 2\beta$, $f_0'(0) = f_0'''(0) = 0$. Both
$\alpha$ and $\beta$ are undetermined parameters.
For the large-$x$ analysis, we impose the physical condition that
$f_0(x) \rightarrow 0$ for $x \to \infty$. This leads to the linearised
version of eq. (\ref{LNSF}) given by
\begin{equation}
\frac{x}{8}\,\frac {df_0}{dx} = \nabla_x^2 \left[\nabla_x^2 f_0 +
\alpha\,f_0 \right]\,.
\label{LIVE}
\end{equation}
There are four linearly independent solutions of (\ref{LIVE}), with the
general asymptotic form
\begin{equation}
f_0(x) \sim F_0 x^c \exp(-Bx^v-Ax^s)\ .
\label{asymp}
\end{equation}
The first solution is the constant solution, corresponding to $A = c = B = 0$.
It satisfies (\ref{LIVE}) by inspection. The other three solutions are
obtained by substituting (\ref{asymp}) into (\ref{LIVE})
and carrying out an asymptotic large-$x$ analysis, leading to the relations
\begin{eqnarray}
v & = & 4/3\,, \nonumber\\
s & = & 2/3 \,,\nonumber\\
B^3 & = & -1/8v^3\,, \nonumber\\
A & = & 64\alpha B^2/9\,, \nonumber\\
c & = & -2d/3\,. \nonumber
\end{eqnarray}
The three different solutions correspond to the three solutions for $B$,
one real, two complex. The real solution, $B = -1/2v = -3/8$, leads to
an exponentially diverging solution for $f_0$:
\begin{equation}
f_0(x) \sim F_0\,x^{-2d/3}\exp(3x^{4/3}/8 - \alpha x^{2/3}),
\end{equation}
while the two complex roots, $B = 3(1 \pm i\sqrt{3})/16$, generate
two solutions which can be combined to give an exponentially decaying
solution with oscillatory behaviour
\begin{equation}
f_0(x) \sim F_0 x^{-2d/3}\exp\left(-\frac{3x^{4/3}}{16} + \frac{\alpha
x^{2/3}
}{2}\right)\cos\left(\frac{3\sqrt{3}x^{4/3}}{16}+
\frac{\alpha\sqrt{3}x^{2/3}}{2} + \varphi_0\right)\,,
\end{equation}
where $F_0$ and $\varphi_0$ are arbitrary constants.
Just as in the Model A case, where $\lambda$ was fixed by imposing
physical conditions on the large-$x$ solution, also in this case we
have an eigenvalue problem in which two parameters $\alpha$ and $\beta$ are
chosen to eliminate the unphysical constant solution and the exponentially
diverging solution. The same problem is encountered in Model B with a
scalar order parameter \cite{MAZ2}. Applying the procedure described in
\cite{FA,MAZ2} it is possible to determine $\alpha$ and $\beta$.
Turning now to the corrections to scaling, we consider first the
four linearly independent large-$x$ solutions for the
linearised form of eq. (\ref{LNCF}). These are a power law solution,
$f_1(x) \sim x^{-\omega}$, an exponentially growing solution, $\sim x^p
\exp(3x^{4/3}/8 - \alpha x^{2/3})$, and two decaying solutions that
can be combined in the form
\begin{equation}
f_1(x)\sim x^p \exp\left(\frac{-3x^{4/3}}{16} +
\frac{\alpha x^{2/3}}{2}\right)
\cos\left(\frac{3\sqrt{3}x^{4/3}}{16}
+ \frac{\alpha \sqrt{3} x^{2/3}}{2} + \varphi_1\right)\,,
\end{equation}
where $\varphi_1$ is arbitrary, and $p = (\omega - 2d)/3$ if $\omega > 4$
and $p = (4 - 2d)/3$ otherwise. The small-$x$ solution
is $f_{1}(x) = \mu x^2 - \alpha\, \mu\, (1+3/n)\,x^4/4(d+2) + \cdots$ .
Therefore (\ref{LNCF}) is solved numerically with initial conditions
$f_1''(0) = 2\mu$, $f_1(0)=f_1'(0)=f_1'''(0)=0$. The two parameters
$\mu$ and $\omega$ are as yet undetermined. They are fixed in the same
way as $\alpha$ and $\beta$, by requiring that an oscillatory,
exponentially decaying solution is recovered as $x\rightarrow\infty$.
Values for $\alpha$, $\beta$, $\mu$ and $\omega$ in 3-$d$ for $n = 2$, 5,
10, 20 and 50 are shown in Table III (For $n = 2$, the effective
$n^* = \pi^2 /4$ has been used). The functions $f_0(x)$ and $f_1(x)$ are
displayed in Figure 2 for $n = 2$ and 20 in 3-$d$. Again $b$ has been set
to $b = 2$ without loss of generality.
The most important result to be extracted for Table III is that the value
of $\omega$ decreases as $n$ increases. This behaviour is quite different
from Model A, where $\omega$ increases asymptotically to $4$ as $n$
increases. It seems from Table III that $\omega$ probably tends to zero for
$n \to \infty$, although an analytical determination of the correction to
scaling for $n$ large but finite, analogous to the treatment of the leading
scaling function in \cite{BH2}, has not yet been realized.
Comparison of the scaling and correction-to-scaling functions displayed in
Figure 2 reinforces a point made in connection with the nonconserved systems,
namely that the correction to scaling become large (relative to the scaling
function itself) at large values of the scaling variable $x$. As noted
before, this suggests that in carrying out scaling analyses of data,
more attention should be paid to small and intermediate values of $x$, where
corrections to scaling can be expected to be (relatively) smaller, than to
large $x$. Indeed, for the nonconserved case (Figure 1) the correction to
scaling has its maximum at a point where the scaling function is already
quite small (around 0.1).
\section{Summary}
Corrections to scaling associated with a non-scaling initial condition have
been studied in $O(n)$ models within the Gaussian closure scheme of Mazenko.
We have calculated both the correction-to-scaling function, $f_1(x)$, and the
associated correction-to-scaling exponent, $\omega$, for both nonconserved and
conserved fields. In both cases Mazenko theory suggests that $\omega$ is
nontrivial, depending on the nature of the dynamics involved, the
dimensionality, $d$, of the system and the number of order parameter
components, $n$. For nonconserved fields the value of $\omega$ tends
to the limiting value $4$ for $n \to \infty$ with $d$ fixed, and for
$d \to \infty$ with $n$ fixed. In the latter limit, the Mazenko theory
reduces \cite{BH1,LiuMaz} to the OJK theory \cite{OJK} and its
generalizations \cite{BPTO}, believed to become exact as
$d \to \infty$ \cite{LIME}.
The 1-$d$ XY model is anomalous in that it exhibits a different growth law
from the standard one for nonconserved dynamics, and the correction-to-scaling
exponent is simple ($\omega = 2$). In this model quantities of interest,
such as the correlation function $C(r,t)$, retain `memory' of the initial
conditions even in the scaling limit.
In studying the conserved $O(n)$ model, an expansion in $1/n$ was used which
is valid for large $n$. This approach was used to find the
correction-to-scaling function $f_1(x)$ and the exponent $\omega$ for
$n = 5$, 20 and 50 in 3-$d$. For $n=2$, an expansion in $C$ up to
$C^3$ was made. In the latter case, a comparison \cite{FA} between
leading-order scaling results and simulations shows very good agreement
despite the wrong growth law (i.e without the logarithmic corrections
predicted for $n=2$ \cite{BR}). In conserved systems $\omega$ decreases
as $n$ increases, raising the question of whether $\omega \rightarrow 0$ or
approaches some limiting value as $n$ becomes very large. We have as yet
been unable to find $f_1(x)$ and $\omega$ analytically in the limit of
large but finite $n$ -- this remains an interesting open question.
The main lesson for the analysis of experimental and simulation data is
that corrections to scaling can be expected to be relatively small at small
and intermediate scaling variable $x$ (=$r/L$), suggesting that this region
be given more weight than large $x$ in fitting (or collapsing) data.
\bigskip
\begin{center}
\begin{small}
{\bf ACKNOWLEDGEMENTS}
\end{small}
\end{center}
This work was supported by EPSRC under grant GR/L97698 (AB) and
by the Commonwealth Scholarship Commission (NR).
|
\section{Introduction}
The asymmetric simple exclusion model \cite{ligget}
is a one-dimensional stochastic model
that describes the time fluctuations of particles diffusing asymmetrically
on the lattice. If we interprete an occupied site as $\sigma_i^z = +1$ and a
vacant site as $\sigma_i^z = -1$ the time evolution of the probability
distribution of particles is given by the following asymmetric XXZ Hamiltonian
\begin{equation}
\label{1}
H = -\sum_{i=1}^N\lbrack\epsilon_+\sigma_j^-\sigma_{j+1}^+ +\epsilon_-
\sigma_j^+\sigma_{j+1}^- + \frac{1}{4}(1-\sigma_j^z\sigma_{j+1}^z)\rbrack,
\end{equation}
where $N$ is the number of lattice sites and $\sigma_i^{\pm} =
(\sigma^x \pm i\sigma^y)/2$ are the raising and lowering Pauli
operators. Periodic boundary conditions are imposed and $\epsilon_+$
and $\epsilon_-$ ($\epsilon_+ + \epsilon_- = 1$) are the transition
probabilities for having a motion to the right and to the left, respectively.
The physical properties of this non-hermitian quantum chain
as well as the related asymmetric six-vertex model are still under
extensive investigations \cite{kim2,albertini1,albertini2}.
This model also describes the surface fluctuations in a growth model
known as single-step model \cite{meakin,liu} where
$(\sigma_i^z + \sigma_{i+1}^z)/2 = -1, 0, 1 $
is the height difference between nearest-neighbor steps located at
the odd-half integers sites $(i-1/2)$ ($i=1,2,\ldots$).
The master equation defining the Hamiltonian \rf{1} can also be interpreted
\cite{spohn}
as the discretized version of the noisy Burges equation or the
Kardar-Parisi-Zhang (KPZ) equation \cite{KPZ}
governing the motion of the height
of growing
surfaces whose local growth velocity depends nonlinearly on the local shape.
The conection between the scaling behavior of the structure function
\cite{liu,krug} of the
stochastic model and the finite-size dependence of the real part of the mass
gap $G_N$ gives us the dynamical critical exponent $z$,
\begin{eqnarray}
\mbox{Re}(G_N) \sim N^{-z}.\nonumber
\end{eqnarray}
This connection enabled Gwa and Spohn \cite{spohn} to explore the Bethe ansatz
solution of
\rf{1} and calculate exactly the exponent $z=3/2$ in the highly anisotropic
limit $\epsilon_- = 0$. Subsequently Kim \cite{kim} extended this result for
$\epsilon_- > 0 $.
In a previous paper \cite{alcabar1} we observed, in connection with a model
for strongly correlated system,
that it is possible to keep the exact integrability of the XXZ chain by
enlarging the excluded volume to the spins. Motivated by these results we
introduce in this paper a generalization of the simple exclusion model
where each particle individually instead of having size 1, in units of
lattice spacing, may have an arbitrary and distinct integer size
$(s_i = 0,1,2,\ldots)$.
Particles with size $s_i > 1 $ will produce stronger excluded volume
than those in the simple exclusion
model where all the particles have unity size $(s_1 = s_2 =\cdots=s_n = 1)$.
Particles with size zero $(s = 0)$
produce no excluded volume since we may put an arbitrary number
of them at the same site. By considering
arbitrary mixtures of molecules with appropriate sizes we may
change continuously the excluded volume in the bulk limit.
The time-evolution operator for these models
are generalizations of ferromagnetic XXZ chain where
restrictions on spin configurations, which depend on the particular
sizes of the molucules $s_i$, are added. We show that for arbitrary
distribution of molecule's sizes the eigenspectrum of the
related Hamiltonian can be calculated exactly through the Bethe
ansatz. The exact integrability for the particular case where all
the molecules have the same size and the diffusion is fully asymmetric
was also verified recently by
Sasamoto and Wadati \cite{sasawada}.
Following Gwa and Spohn \cite{spohn} we show, in the anisotropic limit
$\epsilon_- = 0$, that for arbitrary distribution of molecules the
real part of the gap behaves as $\mbox{Re}(G) \sim N^{-3/2}$, giving a universal
KPZ behaviour with dynamical critical exponent $z=3/2$. Since in our model
the excluded volume can be controlled continuously by
changing the distribution of molecules, the above exact results imply
that the exclusion volume effect is irrelevant to the KPZ dynamics.
The paper is organized as follows. In the next section we introduce the
generalized asymmetric model and the related generalized surface growth
model. The Bethe-ansatz solution of our model is presented in section 3 and
in section 4 a
numerical and analitical calculation of the dynamical critical
exponent $z$ is presented. Finally in section 5 we give our
conclusions and in an appendix we relate exactly
the general asymmetric exclusion model with several boundary conditions with
the simple exclusion model, in different lattice sizes.
\section{The generalized asymmetric exclusion model}
The simplest realization of the model we consider in this paper is
the asymmetric diffusion of molecules (or particles) on a lattice of size $N$,
where each molecules $i$ ($i=1,2,\ldots,n$) may have
a distinct size $s_i = 0,1,2,\ldots$, in units
of lattice spacing.
We represent the molecules on the lattice by placing their center of mass
at the lattice sites. In Fig. 1 we show some examples of configurations
of $n=5$ molecules with size's distribution $\{s\}$ in a lattice with $N=6$
sites.
Molecules of size $s = 0$ are special since
we can put in a given lattice point an
arbitrary number of them. As we can see from Fig. 1, the minimum distance
between two particles with size $s,s'$ on the lattice is given by
\begin{equation}
\label{1-1}
d_{s,s'} = \mbox{Int}\lbrack\frac{s + s' + 1}{2}\rbrack, \;\;\;\;
s,s' = 0,1,2,...,
\end{equation}
where $\mbox{Int}(x)$ is the integer part of $x$.
In order to describe the occupancy of a given
site $i$ ($i=1,2,\ldots,N$)
we attach on it a site variable $\beta_i$ taking integer values
$(\beta_i \in {\rm Z})$. If $\beta_i = 0$ the site is empty, on the other hand
if $\beta_i > 0$ we have a molecule of size $s=\beta_i$ and if
$\beta_i = -n < 0$ we have $n$ molecules of size 0. The allowed
configurations $\lbrace\beta_i\rbrace = \lbrace\beta_1,\beta_2,...,
\beta_N\rbrace$ are those satisfying the constraints imposed by the
sizes of molecules in a periodic lattice, i.e., if $\beta_i \ne 0$ and
$\beta_j \ne 0$ we must have $|i-j|\ge d_{s(\beta_i),s(\beta_j)}$, where
$s(\beta) = 0$ if $\beta
\leq 0$, and $s(\beta) = \beta$ if $\beta > 0$, is the excluded volume
(or size) associated to $\beta$.
The master equation for the probability distribution
$P(\lbrace\beta\rbrace,t)$ can be written in general as
\begin{equation}
\label{2}
\frac{\partial P(\lbrace\beta\rbrace,t)}{\partial t} = - \Gamma(\lbrace\beta\rbrace
\rightarrow \lbrace\beta'\rbrace) P(\lbrace\beta\rbrace,t) +
\Gamma(\lbrace\beta'\rbrace
\rightarrow \lbrace\beta\rbrace) P(\lbrace\beta'\rbrace,t),
\end{equation}
where $\Gamma(\lbrace\beta\rbrace \rightarrow \lbrace\beta'\rbrace)$ is
the transition rate, where a configuration $\lbrace\beta\rbrace$ changes to
$\lbrace\beta'\rbrace$. In the model under consideration we only allow,
whenever it is possible, a single particle to diffuse into its nearest
neighbor sites. The possible motions are diffusion to the right
\begin{eqnarray}
\label{3}
\beta_i\;\;\emptyset_{i+1} &\rightarrow& \emptyset_i\;\;\beta_{i+1},\;\;\;\;\;\;\;\;
\;\;\;\;\;\;\;\beta > 0\nonumber\\
\beta_i\;\;\gamma_{i+1} &\rightarrow& (\beta+1)_i \;\;(\gamma - 1)_{i+1}
,\;\;\;\;\beta < 0,
\gamma \leq 0,
\end{eqnarray}
with transition rate $\epsilon_R$ and diffusion to the left
\begin{eqnarray}
\label{4}
\emptyset_{i}\;\;\beta_{i+1} &\rightarrow&\beta_i\;\;\emptyset_{i+1},\;\;\;
\beta > 0\nonumber\\
\gamma_i\;\;\beta_{i+1} &\rightarrow& (\gamma -1)_i\;\;(\beta+1)_{i+1}
\;\;\;\beta < 0, \;\;\;\; \gamma \leq 0,
\end{eqnarray}
with transition rate $\epsilon_L$. The master equation (2) can be written as
a Schr\"odinger equation in Euclidean time (see ref.\cite{alcrit1} for general
application for two-body processes)
\begin{equation}
\label{5}
\frac{\partial|P>}{\partial t} = -H |P>,
\end{equation}
if we interpret $|P> \equiv P(\lbrace\beta\rbrace, t)$ as the associated wave
function. If we represent $\beta_i$ as $|\beta>_i$ the vector
$|\beta>_1\otimes |\beta>_2 \otimes \cdots \otimes |\beta>_N$ will give us
the associated
Hilbert space. The process \rf{3} and \rf{4} gives us the Hamiltonian
\begin{eqnarray}
\label{6}
H &=& -D \; P\sum_{i=1}^N (H_{i,i+1}^{>} + H_{i,i+1}^{<})P, \nonumber \\
H_{i,j}^{>}& =& \sum_{\beta =1}^{\infty}\lbrack \epsilon_{+}
(E_i^{0,\beta}E_{j}^{\beta,0}- E_{i}^{\beta,\beta}E_{j}^{0,0}) +
\epsilon_{-}(E_i^{\beta,0}E_{j}^{0,\beta }- E_{i}^{0,0} E_{j}^{\beta,\beta})
\rbrack , \\
H_{i,j}^{<}& =& \sum_{\beta =-\infty}^{-1}\sum_{\gamma =-\infty}^{0}\lbrack
\epsilon_{+}
(E_i^{\beta +1,\beta}E_{j}^{\gamma -1,\gamma}- E_{i}^{\beta,\beta}E_{j}^
{\gamma,\gamma}) +
\epsilon_{-}(E_i^{\gamma-1,\gamma}E_{j}^{\beta +1,\beta }-
E_{i}^{\gamma,\gamma} E_{j}^{\beta,\beta})
\rbrack \nonumber ,
\end{eqnarray}
with
\begin{equation}
\label{7}
D = \epsilon_R + \epsilon_L,\;\;\; \epsilon_+ =\frac{\epsilon_R}{\epsilon_R +
\epsilon_L},\;\;\; \epsilon_- = \frac{\epsilon_L}{\epsilon_R + \epsilon_L},
\end{equation}
and periodic boundary conditions. The matrices $E^{\alpha,\beta}$ are
infinite-dimensional with a single non-zero
element $(E^{\alpha,\beta})_{i,j} = \delta_{\alpha,i}\delta_{\beta,j}
(\alpha,\beta,i,j \in {\rm Z})$. The projector P keeps on the Hilbert
space only the vectors $|\lbrace\beta\rbrace>$
satisfying the constraint \rf{1-1} which mathematically means that
for the all
$\beta_i,\beta_j \ne 0,\;\; |i - j| \ge d_{s(\beta_i),s(\beta_j)}$.
The constant D in \rf{6} fixes the time scale and for simplicity we
chose $D = 1$. A particular simplification of the above Hamiltonian
occurs when all the molecules have the same fixed size $s > 0$. In this case
the Hamiltonian can be expressed in terms of spin-$\frac{1}{2}$ Pauli
matrices
\begin{equation}
\label{8}
H_{\lbrace s_1=\cdots=s_n = s\rbrace} = - \;
P_s\{\sum_{i=1}^N\lbrack \epsilon_+
\sigma_i^-\sigma_{i+1}^+ +\epsilon_-\sigma_i^+\sigma_{i+1}^- +
\frac{1}{2}(\epsilon_+ + \epsilon_-)(\sigma_i^z\sigma_{i+1}^z - 1)\rbrack\} P_s,
\end{equation}
where now $P_s$ projects out configurations where two up spins, in
$ \sigma^z$ basis, are at distance smaller than the size $s > 0$. In the
case where $s=1$ the projector $P_s = 1$ and we have the standard asymmetric
exclusion Hamiltonian \rf{1}. In terms of Pauli matrices this operator
has the general form
\begin{equation}
\label{9}
P_s = \prod_{i}\Bigl[\frac{1}{2}(1 - \sigma_i^z) +
\frac{1}{2}(1 + \sigma_i^z)
\prod_{l=1}^{s-1}(\frac{1 - \sigma_{i+l}^z}{2})\Bigr].
\end{equation}
The Hamiltonian \rf{8} can be more easily compared with standard magnetic
quantum chains by performing for $\epsilon_+,\epsilon_- \ne 0$ the following
canonical change of variables
\begin{equation}
\label{10}
\sigma_i^{\pm} \rightarrow (\frac{\epsilon_-}{\epsilon_+})^{\pm i/2}
\sigma_i^{\pm},\;\;\; \sigma^z \rightarrow \sigma^z,\;\;\; i = 1,2,...,N,
\end{equation}
which gives
\begin{eqnarray}
\label{11}
H &=& -\frac{1}{2}\sqrt{\epsilon_+\epsilon_-}\sum_{i=1}^N P_s\Bigl[
\sigma_i^x\sigma_{i+1}^x
+ \sigma_i^y\sigma_{i+1}^y + \Delta (\sigma_i^z\sigma_{i+1}^z-1)\Bigr]
P_s,\nonumber\\
\Delta &=&\frac{\epsilon_+ + \epsilon_-}{\sqrt{\epsilon_R \epsilon_L}}.
\end{eqnarray}
Apart from the projector $P_s$ this is the ferromagnetic XXZ chain or
the anisotropic Heisenberg chain. However in distinction with (8) the boundary
conditions are now twisted
\begin{equation}
\label{12}
\sigma_{N+1}^{\pm} = (\frac{\epsilon_+}{\epsilon_-})^{\pm N/2}
\sigma_1^{\pm},\;\;\; \sigma_{N+1}^z =\sigma_1^z.
\end{equation}
We expect that ferromagnetic quantum chains like those in \rf{11} are gapped
for $\Delta > 1$ However since
$(\epsilon_+/\epsilon_-)^{N/2}\rightarrow \infty$, for
$N\rightarrow \infty$ the boundary condition gives us interaction with the
same degree of importance as the totality of the other interactions (see
\cite{henkel} for a related problem). As we
will see, from the exact solution of \rf{6} and \rf{8}, this surface term is
strong enough to produce a gapless eigenspectrum.
In surface growth physics the asymmetric
simple exclusion model is related
to the single-step model. Similarly our generalized model is also
related to a generalization of the single-step model. The surface
configurations in this growth model are obtained by defining
height variables $h_i$ $(i =1,2,\ldots )$ which are related to the
spin variables in
our generalized asymmetric diffusion model. For simplicity we are going to
present only the
surface growth model related to the diffusion problem where
all molecules have the
same size $s$. Let us consider initially $s > 0$. For a given
configuration $\lbrace\beta_1,...,\beta_N\rbrace$ of molecules of size $s$,
the height variables should obey
\begin{equation}
\label{f1}
h_{i+1} - h_i = f(\beta_{i -\frac{s}{2}}, \beta_{i -\frac{s}{2}+1},...,
\beta_{i +\frac{s}{2}}, \beta_{i +\frac{s}{2}+1}),
\end{equation}
where $f = 0$ for all allowed configurations except in the case
\begin{equation}
\label{f2}
f(0,\ldots,0) = -1\;\;\;\; \mbox{and}\;\;\;\; f(1,0,\ldots,0,1) = 1.
\end{equation}
The variables $\lbrace\beta\rbrace$ of the related diffusion model are defined
at the links or at the same
positions of the height variables $\lbrace h \rbrace$
depending if the size of the molecules $s$ is odd or even, respectively.
The number of molecules $n$ $(0,1,\ldots)$,
in the generalized asymmetric diffusion
is
conserved and for each value of $n$ we are going to have a growth model with
different boundary
conditions in the spacial direction. The dynamical
rules defining the growth are the following
a) No steps on the surface are
allowed to be higger than 1, in units of lattice spacing in the growth
direction, i. e.,
\begin{equation}
\label{f3}
h_{i+1} -h_i = 1, 0,-1 \quad (i =1,\ldots,n-1).
\end{equation}
b) All the local valleys and hills should have a size, in units
of the lattice spacing in the spatial direction, which is a multiple of
$b = s+1$.
c) The following boundary condition should be satisfied
\begin{equation}
\label{f4}
h_{i+N} = h_i - \bar{h}, \;\;\;\;\bar{h} = b\lbrace[N/b]_I - n\rbrace
+ [N/b]_R .
\end{equation}
where by $[N/b]_I$ and $[N/b]_R$ we mean the integer part and the rest
of the division $N/b$.
d) The surface changes whenever obeying the previous requirements we
can still adsorb $ (h_{i+l} \rightarrow h_{i+l}+1, l = 0,1,...,b-1)$
or desorb $(h_{i+l}\rightarrow h_{i+l}-1, l =0,1,...,b-1)$ at height
$h_i$ $(i=1,\ldots,N)$ a retangular brick of size $b$ in the spatial
direction and size 1 in the growth direction.
We choose a height $h_i$ $ (i =
1,...,n)$ at random. If it is possible to adsorb or desorb a brick,
with probability $\epsilon_+/2$ $ (\epsilon_-/2)$ we desorb (adsorb)
a brick, and do nothing with probability $\frac{1}{2}$. If it is possible
, at $h_i$, only to desorb (adsorb) a brick, with probability
$\epsilon_+$ $(\epsilon_-)$ we desorb (adsorb) a brick and
with probability $1-\epsilon_+$ $ (1 - \epsilon_-)$ we do nothing. In Fig. 2
we
show for $N = 7, s = 2$ $ (b = 3)$ and $n=2$ some examples of the
possible configurations of the surface. In this figure we also
show the corresponding particle configurations in the diffusion
problem. We can verify that for arbitrary $s $ (or b), as
long the growth model is not periodic $(N \ne n b)$ there exists
an exact
one-to-one correspondence between the configurations of particles and
those of the surface, with the transitions between them described
by the Hamiltonian \rf{6}. On the other hand, if the growth model
is periodic $(N = n b)$, there exist $b$ configurations in the
asymmetric diffusion problem that correspond to a single surface
configuration (the flat surface). Consequently the Hamiltonian
\rf{6} does not describe exactly the generalized step
model in a finite lattice.
However as $N$ increases this difference decreases and \rf{6} also
describes the fluctuations of the growth model.
Finally, in the case where all the molecules has size zero
a possible growth model is obtained by defining height variable
$h_i$ $ (i=1,2,...,N)$ at the same position of the molecules in
the diffusion problem. For a given configuration
$\lbrace n_1,n_2,...,n_N\rbrace$ with $n_i$ molecules at sites $i$
the height variables in the surface model $(h_{i+1} \geq h_{i})$
satisfy
\begin{equation}
\label{f5}
h_i -h_{i-1} = n_i,\;\;\;\; i = 2,...,N,
\end{equation}
with the boundary condition
\begin{equation}
\label{f6}
h_{N+1} = h_1 + n
\end{equation}
where $n=\sum n_i$ is the total number of molecules. Bricks of
unity size are desorbed (adsorbed) with transition rates
$\epsilon_+$ $(\epsilon_-)$ if the final configuration satisfies
$h_{i+1}\geq h_i$ $ (i=1,2,\ldots,N-1)$.
\section{ The Bethe ansatz equations}
We present in this section the exact solution of the general quantum chain
\rf{6}. For simplicity let us consider initially the case where all the
molecules have the same size
$s$ ($0,1 ,\ldots$). In the particular case
where $s = 1$ we have the standard simple exclusion model whose Bethe
ansatz solution was obtained by Gwa and Spohn \cite{spohn} and can also be
obtained after the canonical transformation \rf{12} from the Bethe ansatz
solution of the XXZ chain with twisted boundary conditions \cite{alcbat1}.
The exact integrability of the fully asymmetric version of \rf{11}
($\epsilon_- = 0$), for $s >0$, was verified directly in the master equation
by Sasamoto and Wadati \cite{sasawada}, and the model with $s =0$ is related to
the limit $q \rightarrow \infty $ of the $q$-boson hopping model introduced
by Bogoliubov {\it et al}. \cite{bogo1,bogo2}.
Due to the conservation of particles in the diffusion processes the
total number of particles are good quantum numbers and we can separate
the associated Hilbert space into block-disjoint sectors labelled by the
number $n$ of particles. We therefore consider the eigenvalue equation
\begin{equation}
\label{13}
H |n> = E|n>,
\end{equation}
where
\begin{equation}
\label{14}
|n> = \sum_{\{x\}}f(x_1,\ldots,x_n)|x_1,\ldots,x_n>.
\end{equation}
Here $x_1,\ldots,x_n$ denotes the location of particles on the chain and the
summation extends over all sets $\{x\}$ of the $n$ non-decreasing
integers satisfying
\begin{equation}
\label{15}
x_i \geq x_{i+1} + s, \quad i=1,\ldots,n-1, \quad s \leq x_n -x_1 \leq N-s.
\end{equation}
It is important to notice that some of
these coordinates may coincide in the case where the particles have
zero size.
{\it {\bf n = 1.}}
For one particle on the chain as a consequence of the translational invariance
of \rf{6} the eigenfunctions are the momentum-$k$ eigenfunctions
\begin{equation}
\label{16}
|1> = \sum_{x=1}^{N} f(x) |x>, \; \; f(x) = e^{ikx},
\end{equation}
where
\begin{equation}
\label{16-k}
k =\frac{2\pi}{N}l; \quad l=0,1, \ldots,N-1,
\end{equation}
and energy given by
\begin{equation}
\label{17}
E = e(k) \equiv -(\epsilon_-e^{ik} + \epsilon_+e^{-ik} -1).
\end{equation}
{\it {\bf n =2.}} For two particles on the lattice the eigenvalue equation
\rf{13} gives us two distinct relations depending on the relative position of
the particles. If the two particles are at positions $x_1$ and $x_2$
satisfying $x_2 >x_1 +s$ we obtain
\begin{eqnarray}
\label{18}
Ef(x_1,x_2) &=& -\epsilon_+f(x_1-1,x_2) - \epsilon_-f(x_1+1,x_2)
+ 2f(x_1,x_2) \nonumber \\
& & -\epsilon_+f(x_1,x_2-1) - \epsilon_-f(x_1,x_2+1),
\end{eqnarray}
that can be solved by the ansatz
\begin{equation}
\label{19}
f(x_1,x_2) = e^{ik_1x_1}e^{ik_2x_2},
\end{equation}
which gives
\begin{equation}
\label{20}
E = e(k_1) + e(k_2).
\end{equation}
Since the relation \rf{20} is symmetric in $k_1$ and $k_2$ we can write
a more general solution of \rf{18} as
\begin{equation}
\label{21}
f(x_1,x_2) = A_{12}e^{ik_1x_1}e^{ik_2x_2} + A_{21}e^{ik_2x_1}e^{ik_1x_2}
\end{equation}
with the same energy as in \rf{20}. The second relation happens when
$x_2 = x_1 + s$. In this case we have
\begin{equation}
\label{22}
Ef(x_1,x_1 +s) = -\epsilon_+f(x_1-1,x_1+s)-\epsilon_-f(x_1,x_1+s+1)
+ f(x_1,x_1+s) .
\end{equation}
If we now substitute the ansatz \rf{21} with the energy \rf{20}, the
constants $A_{12}$ and $A_{21}$, initially arbitrary, should now satisfy
\begin{equation}
\label{23}
\frac{A_{12}}{A_{21}}= - \Bigl(\frac{e^{ik_1}}{e^{ik_2}}\Bigr)
^{s-1}e^{i\Psi_{12}},
\end{equation}
\begin{equation}
\label{24}
e^{i\Psi_{jl}} =\frac{\epsilon_+ + \epsilon_-e^{i(k_j+k_l)} - e^{ik_j}}
{\epsilon_+ + \epsilon_-e^{i(k_j+k_l)} - e^{ik_l}}.
\end{equation}
The "wave numbers" $k_1$ and $k_2$ are complex in general and
are fixed due to the cyclic boundary
condition
\begin{equation}
\label{25}
f(x_2,x_1+N) = f(x_1,x_2),
\end{equation}
which from \rf{21} give us the equations
\begin{equation}
\label{26}
\frac{A_{12}}{A_{21}} = e^{-ik_2N}, \quad \frac{A_{21}}{A_{12}} = e^{-ik_1N}.
\end{equation}
Equations \rf{23} and \rf{24} give us the Bethe-ansatz equations for
$n =2$
\begin{equation}
\label{27}
e^{ik_jN} = -\prod_{l=1}^2 \Bigl( \frac{e^{ik_j}}{e^{ik_l}}
\Bigr)^{s-1} e^{i\Psi_{j,l}},
\; \; j=1,2 \;\;,
\end{equation}
with energy given by \rf{20}.
{\it {\bf General n.}} The above calculation can be generalized for arbitrary
values of $n$. The ansatz for the wavefunction becomes
\begin{equation}
\label{28}
f(x_1,\ldots,x_n) = \sum_P A_{P_1,P_2,\ldots,P_n} e^{i(k_{P_1}x_1+\cdots +
k_{P_n}x_n)},
\end{equation}
where the sum extends over all permutations $P$ of $1,2,\ldots,n$. If
$x_{i+1} -x_i >s$ for $i=1,2,\ldots,n$ it is easy to see
that the eigenvalue equation \rf{13} is satisfied by the ansatz \rf{28}
with energy
\begin{equation}
\label{29}
E = \sum_{j=1}^n e(k_j).
\end{equation}
If a pair of paticles is at positions $x_i,\; x_{i+1}$, where $x_{i+1} =
x_i +s$, equation \rf{13} with the ansatz \rf{28} and the relation
\rf{29} give us
\begin{equation}
\label{30}
\frac{A_{P1,\ldots,P_i,P_{i+1},\ldots,,P_n}}
{A_{P_1,\ldots,P_{i+1},P_i,\ldots,P_n}} = -e^{i(s-1)(k_{P_i} -k_{P_{i+1}})}
e^{i\Psi_{P_i,P_{i+1}}}.
\end{equation}
Inserting the ansatz \rf{28} in the boundary condition
\begin{equation}
\label{31}
f(x_2,\ldots,x_n,x_1 + N) = f(x_1,\ldots,x_n)
\end{equation}
we obtain the additional relation
\begin{equation}
\label{32}
A_{P_1,\ldots,P_n} = e^{ik_{P_1}N}A_{P_2,\ldots,P_n,P_1}.
\end{equation}
If we iterate the relation \rf{30} $n$ times the equation \rf{32} gives us the
Bethe-ansatz equations
\begin{equation}
\label{33}
e^{ik_jN} = (-)^{n-1} \prod_{l=1}^n \Bigl(\frac{e^{ik_j}}{e^{ik_l}}
\Bigr)^{s-1}
\frac{\epsilon_+ + \epsilon_- e^{i(k_j+k_l)} -e^{ik_j}}
{\epsilon_+ + \epsilon_- e^{i(k_j+k_l)} -e^{ik_l}}
\end{equation}
for $j=1,2,\ldots,n$. The solutions $\{k_j\}$ of these equations with
\rf{29} give us the eigenergies of the Hamiltonian \rf{6}. Furthermore,
it follows from a lattice shifting that the wavefunctions given by the
ansatz \rf{28} are also eigenfunctions of the momentum operator $\hat{P}$
with eigenvalue
\begin{equation}
\label{34}
p = \sum_{j=1}^n k_j \quad \mbox{mod} (2\pi) = \frac{2\pi}{N}l,
\quad l=0,1,\ldots,
N-1.
\end{equation}
In the particular case where $s=1$ equations \rf{29},\rf{33} and \rf{34}
recover the results presented in Ref. \cite{spohn}.
Let us now consider the general case where we have $n$ molecules with
arbitrary given sizes $\{s_1,s_2,\ldots,s_n\}$ ($s_i=0,1,2,\ldots$) and the
Hamiltonian given by \rf{6}.
In this case each type of molecule is conserved
separately. Moreover since in the diffusion processes the particles only
interchange positions with the vacant sites a given order
$\{s_1,s_2,\ldots,s_n\}$ of particles remain conserved up to cyclic
permutations. The wavefunctions can be written as
\begin{equation}
\label{37}
|s_1,s_2,\ldots,s_n> = \sum_{\{c\}} \sum_{\{x\}}
f^{s_{c_1},\ldots,s_{c_n}}(x_1,\ldots,x_n)|x_1,\ldots,x_n>.
\end{equation}
Here $f^{s_{c_1},\ldots,s_{c_n}}(x_1,\ldots,x_n)$ is the amplitude of a
configuration where particles of sizes $s_1,\ldots,s_n$ occupy the
positions $x_1,\ldots,x_n$, respectively. The summation $\{c\}$ extends over
all cyclic permutations $\{c_1,\ldots,c_n\}$ of integers $\{1,\ldots,n\}$,
and the summation $\{x\}$ extends, for a given distribution
$\{s_{c_1},\ldots,s_{c_n}\}$ of molecules, to increasing integers satisfying
\begin{equation}
\label{38}
x_{i+1} \geq x_i + d_{s_{c_i},s_{c_{i+1}}} \quad
i =1,\ldots,n-1, \;\; d_{s_{c_n},s_{c_1}} \leq x_n -x_1 \leq
N- d_{s_{c_n},s_{c_1} } .
\end{equation}
The ansatz we expect to be valid, that replaces \rf{27} is
\begin{equation}
\label{39}
f^{s_1,\ldots,s_n}(x_1,\ldots,x_n) = \sum_P A_{P_1,\ldots,P_n}
^{s_1,\ldots,s_n} e^{i(k_{P_1} + \cdots + k_{P_n})}
\end{equation}
where $A_{k_{P_1},\ldots,k_{P_n}}^{s_1,\ldots,s_n}$ and $\{k_1,\ldots,k_n\}$
are going to be fixed by imposing that \rf{38} with \rf{39} is a solution
of the eigenvalue equation of the general Hamiltonian \rf{6}.
Let us consider the eigenvectors of \rf{6} with different number of
particles.
{\it \bf{n = 1.}} For one particle on the chain the ansatz \rf{39} coincides
with \rf{28} and the wavefunctions and energies are given by \rf{16}
and \rf{17}, respectively.
{\it \bf {n = 2.}} If both particles are identical $s_1 = s_2=s $,
we have the same situation considered previously in \rf{18}-\rf{27}.The
wavefunctions $|s,s>$ are given by \rf{39} with energies given by
\rf{20} and \rf{27}. However, if the particles are distinct the situation
is different. If the particles are located at positions $x_1$ and $x_2$,
with $x_2 -x_1 > d_{s_1,s_2}$ the ansatz \rf{38} is valid with energy
given by \rf{20} and no restrictions on $\{A_{k_{P_1},k_{P_2}}^{\alpha_i,
\alpha_j}\}$ are necessary.
If the particles are at the closest distance $x_2 = x_1 +
d_{s_1,s_2}$ equation \rf{22} is replaced by
\begin{eqnarray}
\label{40}
Ef^{s_1,s_2}(x_1,x_1 + d_{s_1,s_2})& =& -\epsilon_+f^{s_1,s_2}(x_1-1,x_1 +
d_{s_1,s_2}) -\epsilon_-f^{s_1,s_2}(x_1,x_1 +d_{s_1,s_2} +1) \nonumber \\
& & +f^{s_1,s_2}(x_1,x_1 + d_{s_1,s_2}).
\end{eqnarray}
Inserting in the above equation the ansatz \rf{39} and the energy \rf{20}
we obtain the relation
\begin{equation}
\label{41}
A_{P_1,P_2}^{s_1,s_2} = -e^{i\Psi_{P_1,P_2}} \
\sum_{s_1',s_2'} S_{s_2',s_1'}^{s_1,s_2}
(k_{P_1} -k_{P_2}) A_{P_2,P_1}^{s_1',s_2'}
\end{equation}
where $\Psi_{P_1,P_2}$ is given by \rf{24} and the elements of the $S$
matrix are given by
\begin{equation}
\label{42}
S_{\gamma,\mu}^{\alpha,\beta}(k) = e^{i(d_{\alpha,\beta}-1)k} \delta_{\alpha,
\mu} \delta_{\beta,\gamma}.
\end{equation}
The wave
numbers $k_1$ and $k_2$ are going to be fixed by the boundary condition
\begin{equation}
\label{43}
f^{s_2,s_1}(x_2,x_1+N) = f^{s_1,s_2}(x_1,x_2),
\end{equation}
but instead of deriving the Bethe-ansatz equations for $n=2$ let us consider
the case of general $n$.
{\it {\bf General n}}. The ansatz \rf{39} applied to the
case where two particles are at their closest distance gives us the
generalization of \rf{41}
\begin{equation}
\label{44}
A_{\ldots,P_1,P_2,\ldots}^{\ldots,\alpha,\beta,\ldots} = -e^{i\Psi_{P_1,P_2}}
\sum_{\alpha',\beta'} S_{\alpha',\beta'}^{\alpha,\beta}
(k_{P_1}-k_{P_2}) A_{\ldots,P_2,P_1,\ldots}^{\ldots,\beta',\alpha',\ldots},
\end{equation}
with $S$ given by \rf{42}. Successive applications of this equation give us in
general different relations between the amplitudes. For example $A_{\ldots,
k_1,k_2,k_3,\ldots}^{\ldots,\alpha,\beta,\gamma,\ldots}$ relate to
$A_{\ldots,k_3,k_2,k_1,\ldots}^{\ldots,\gamma,\beta,\alpha,\ldots}$ by
performing the permutations $\alpha\beta\gamma \rightarrow \beta\alpha\gamma
\rightarrow \beta\gamma\alpha \rightarrow \gamma\beta\alpha$ or
$\alpha\beta\gamma \rightarrow \alpha\gamma\beta
\rightarrow \gamma\alpha\beta \rightarrow \gamma\beta\alpha$, and consequently
the $S$-matrix should satisfy the Yang-Baxter \cite{yang,baxter} equation
\begin{eqnarray}
\label{45}
\sum_{\gamma,\gamma',\gamma''} S_{\gamma,\gamma'}^{\alpha,\alpha'}(k_1-k_2)
S_{\beta,\gamma''}^{\gamma,\alpha''}(k_1-k_3)& &S_{\beta',\beta''}^
{\gamma',\gamma''}(k_2-k_3) = \nonumber \\
& & \sum_{\gamma,\gamma',\gamma''}
S_{\gamma',\gamma''}^{\alpha',\alpha''}(k_2-k_3)
S_{\gamma,\beta''}^{\alpha,\gamma''}(k_1-k_3)S_{\beta,\beta'}^
{\gamma,\gamma'}(k_1-k_2).
\end{eqnarray}
Actually the relation \rf{45}
is a necessary and sufficient condition \cite{yang,baxter}
to obtain non-trivial
solution for the amplitudes in \rf{44}. The validity of \rf{45} can easily
be verified for the diagonal $S$-matrix \rf{42} in the problem we are
considering.
The boundary condition
\begin{equation}
\label{46}
f^{s_1,\ldots,s_n}(x_1,\ldots,x_n) = f^{s_2,\ldots,s_n,s_1}(x_2,\ldots,x_n,x_1)
\end{equation}
implies the relation between the amplitudes
\begin{equation}
\label{47}
A_{P_1,\ldots,P_n}^{s_1,\ldots,s_n} = e^{ik_{P_1}N}
A_{P_2,\ldots,P_n,P_1}^{s_2,\ldots,s_n,s_1}.
\end{equation}
If we now apply relation \rf{44} $n$ times we can obtain a relation between
the amplitudes with same momenta, i. e.,
\begin{eqnarray}
\label{48}
& & A_{P_1,\ldots,P_n}^{s_1,\ldots,s_n} = (-)^{n-1} e^{i\sum_{l=2}^n
\Psi_{P_l,P_1}} e^{ik_{P_1}N} \nonumber \\
& & \sum_{\{s_1',\ldots,s_n'\}} \sum_{\{s_1'',\ldots,s_n''\}}
S_{s_1',s_1''}^{s_1,s_2''}(k_{P_1}-k_{P_1})
S_{s_2',s_2''}^{s_2,s_3''}(k_{P_2}-k_{P_1}) \cdots
S_{s_n',s_n''}^{s_n,s_1''}(k_{P_n}-k_{P_1}) A_{P_1,\ldots,P_n}^
{s_1',\ldots,s_n'},
\end{eqnarray}
where we introduced the extra sum
\begin{equation}
\label{49}
1 = \sum_{s_2'',s_1''} \delta_{s_2'',s_1'} \delta_{s_1,s_1''} =
\sum_{s_2'',s_1''} S_{s_1',s_1''}^{s_1,s_2''}(k_{P_1}-k_{P_1}).
\end{equation}
In order to fix the values of $\{k_j\}$ we should then find the eigenvalues
$\Lambda (k)$ of the matrix
\begin{equation}
\label{50}
<\{s\}|T(k)|\{s'\}> = \sum_{\{s_1'',\ldots,s_n''\}} \prod_{l=1}^n
S_{s_l',s_l''}^{s_l,s_{l+1}''}(k_{P_l} -k),
\end{equation}
with $s_{n+1}'' = s_1''$.
We identify $T(k)$ as the transfer matrix of a inhomogeneous vertex model,
with inhomogeneities $(k_{P_l} - k)$, in a periodic lattice.
The Boltzmann weights of the vertex models are given by $S_{\gamma \delta}
^{\alpha \beta}$ and the number of distinct vertices depends on the particular
configuration (type of order) of molecules in our diffusive system.
Using \rf{42} we can see that there exists only one non-zero element
for each line, i. e., $<s_1,\ldots,s_n|T|s_2,\ldots,s_n,s_1>$.
In order to calculate the eigenvalues $\Lambda(k)$ we apply the transfer
matrix $r$ times in the state $A^{s_1,\ldots,s_n}$, where $r$ is the
minimum number of cyclic rotations of $\{s_1,\ldots,s_n\}$ where the
configuration repeats the initial one. We may show that
\begin{equation}
\label{51}
T^rA_{P_1,\ldots,P_n}^{\alpha_1,\ldots,\alpha_n} = \Lambda^r(k) A_{P_1,\ldots,P_n}^{\alpha_1,\ldots,\alpha_n}.
\end{equation}
Also it is easy to compute
\begin{equation}
\label{52}
\Lambda^r(k) =\exp \Bigl[ i\frac{r}{n} (\sum_{l=1}^n d_{s_l,s_{l+1}}-1)
\sum_{l=1}^n (k_l-k) \Bigr].
\end{equation}
Finally, substituting $\Lambda(k_{P_1})$ in \rf{48} we obtain the
Bethe-ansatz equations
\begin{equation}
\label{53}
e^{ik_jN} = e^{i\frac{2\pi}{r}m} (-1)^{n-1} \prod_{l=1}^n
e^{i(k_j-k_l)(\tilde s-1)} \frac{\epsilon_+ + \epsilon_- e^{i(k_j+k_l)} -
e^{ik_j}}{\epsilon_+ + \epsilon_-e^{i(k_j+k_l)} - e^{ik_l}},
\end{equation}
where $j=1,\ldots,n$; $m=0,1,\ldots,r-1$ and
\begin{equation}
\label{54}
\tilde s = \frac{1}{n} \sum_{l=1}^n d_{s_l,s_{l+1}}
\end{equation}
plays the role of an average molecule size of the particular
configuration $\{s\}$ of molecules. As we can see, by comparing \rf{53} with
\rf{33} the extra phase $\exp(i2\pi m/r)$ ($m =0,1,\ldots,r-1$) gives $r$ times
more solutions of \rf{53} than in \rf{33}. This indeed should be the case
since the Hilbert space associated to the Hamiltonian \rf{6} of a given
distribution of particles of sizes $\{s_1,\ldots,s_n\}$, due to the
distinguibility of the particles, is $r$ times bigger than that of the
Hamiltonian \rf{6} when $s_1=s_2=\cdots=s_n$. It is interesting to observe
that $\tilde s$ can take any non-negative
rational number by choosing appropriately
$\{s_1,\ldots,s_n\}$. Also many distinct distributions of molecule's sizes
with the same effective $\tilde s$ will have the same eigenenergies.
In an appendix we explore our Bethe-ansatz solution to obtain the relationship
between the eigenvalues and eigenvectors of the Hamiltonian \rf{6} with
different distributions of molecule's sizes.
\section{The critical exponent z}
In this section we calculate the dynamical critical exponent $z$ for the
stochastic models presented in Sec. 2. This calculation is done by exploiting
its connection with the mass gap $G_N$ of the corresponding Hamiltonian,
\begin{equation}
\label{55}
\mbox{Re}(G_N) \sim N^{-z}.
\end{equation}
A calculation for arbitrary values of $\epsilon_+,\epsilon_-$ and density
$n/N$ can be done systematically by using the method presented in \cite{kim}.
However, since universality arguments indicate that $z$ should be independent
on the particular values of $\epsilon_+, \epsilon_-$ and $n$, as long as
$\epsilon_+\neq \epsilon_-$, we are going to restrict ourselves, like in
\cite{spohn}, to the simplest case where $\epsilon_- =0, \epsilon_+ =1$. A
general discussion for the other cases, which does not change our results
is given at the end of the appendix. Defining the variables
\begin{equation}
\label{56}
z_j = 2e^{-ik_j} -1
\end{equation}
the energies \rf{29} and momenta \rf{34} are given by
\begin{equation}
\label{57}
E = \sum_{j=1}^n (1 - z_j)/2,
\end{equation}
\begin{equation}
\label{58}
e^{-iP} = \prod_{j=1}^n (1 + z_j)/2,
\end{equation}
respectively. The $\{z_j\}$ variables should satisfy the Bethe-ansatz
equations \rf{53}
\begin{equation}
\label{59}
(1+z_j)^{N-n\tilde s} (1 - z_j)^n = -2^N e^{-i\frac{2\pi}{r} m}
\prod_{l=1}^n \frac{z_l -1}{(z_l+1)^{\tilde s}}
\end{equation}
where $j=1,\ldots,n$, $m=0,1,\ldots,r-1$. It is interesting to note that
these equations are simpler than the usual Bethe-ansatz equations appearing
in other exact integrable systems since the right-hand side of \rf{59} is
independent of the particular value of $j$. These equations are even
simpler for the special "half-filled" density
$\rho = n/N = 1/(1+\tilde s)$, i. e.,
\begin{eqnarray}
\label{60}
(1 - z^2)^n = Y \nonumber \\
Y = -2^{n(1 + \tilde s)} \prod_{l=1}^n \frac{z_l -1}{(z_l+1)^{\tilde s}}.
\end{eqnarray}
If we parametrize $Y = -a^n e^{in \theta }$, with $a \geq 0$ and
$\theta \in (-\frac{\pi}{n},\frac{\pi}{n})$, the $2n$ roots $z_j$ are
given by
\begin{eqnarray}
\label{61}
z_j &=& (1 - y_j)^{\frac{1}{2}},
\quad z_{j+n} = - z_j; \nonumber \\
y_j &=& a e^{i\theta} e^{i2\pi(j-\frac{1}{2})/n}; \quad j =1,\ldots,n.
\end{eqnarray}
For a given choice $\{z_j\}$ of the above set and a given value of $m$
($0,1,\ldots,r-1$) we have only two unknowns, $a$ and $\theta$, which are
obtained from the equation
\begin{equation}
\label{62}
(ae^{i\theta})^n = e^{i\frac{2\pi}{r}m} 2^{(\tilde s +1)n}
\prod_{l=1}^n \frac{z_{l(j)} -1}{[z_{l(j)} +1]^{\tilde s}}.
\end{equation}
We have solved numerically the above equations for several values of
$\tilde s, m, r$ and $n$. The ground-state energy $E=0$ is
obtained by choosing $m=0$ in \rf{62}, and is given by the
configuration
\begin{equation}
\label{63}
C_0 = \{z_1,z_2,\ldots,z_n\},
\end{equation}
with $a = \theta = 0$. In order to find the first excited state we should
consider different choices of $\{z_j\}$ and different values of $m$.
Since $z_j + z_{n+j} = 0$ the energy increases as we take, in the
configurations $\{z_j\}$, values of $z_j$ where $n<j\leq 2n$.
Therefore configurations $\{z\}$ associated with
low energies are
\begin{equation}
\label{64}
C_1 = \{z_1,z_2,\ldots,z_{n-1},z_{n+1}\}
\end{equation}
and
\begin{equation}
\label{65}
C_{-1} = \{z_{2},\ldots,z_{n-1},z_n,z_{2n}\}.
\end{equation}
These configurations, from \rf{58}, correspond to states with momentum
$-\frac{2\pi}{N}$ and $\frac{2\pi}{N}$, respectively. Our numerical
results show that the energy corresponding to the configuration $C_0$ with
$m\neq 0$ behaves for large $N$ as
\begin{equation}
\label{66}
E_{C_0,m} \sim
\frac{a}{n^{\frac{1}{2}}} - i\frac{\pi}{r(\tilde s +1)}m \quad
m = 1, 2, \ldots \;\;.
\end{equation}
On the other hand, the configuration $C_1$ or $C_{-1}$, for sufficiently
large values of $N$ produces the lowest energy when $m=0$, independently of
$\tilde s, r$ and behaves as
\begin{equation}
\label{67}
E_{C_{\pm 1},0} \sim \frac{a_0}{n^z} + i\frac{b_0}{n^\gamma}, \; \;
z = \frac{3}{2}, \;\; \gamma =1,
\end{equation}
where $a_0$ and $b_0$ are constants. The energies for different values of $m$
but with configurations $C_{\pm 1}$ also behave similarly as \rf{67}.
Comparing \rf{67} with \rf{66} we see that the gap is given by
$E_{C_{\pm 1},0}$ and is real only for the special case $\tilde s =1$, treated in \cite{spohn}.
The values of $a$ and $\theta$ that correspond to the first excited state
behaves asymptotically as
\begin{equation}
\label{68}
a = 1 + \frac{\beta}{n} + o(n^{-1}), \;\; \theta = \pm \alpha\frac{
(\tilde s -1)} {n^{\frac{3}{2}}} + o(n^{-\frac{3}{2}}),
\end{equation}
where $\beta$ and $\alpha$ are constants.
In order to illustrate our numerical results we show in Table 1 the
finite-size estimates for the amplitudes $a_0, b_0$ and the exponents $z$ and
$\gamma$ defined in equation \rf{67}.
Accepting the behavior \rf{68} for the values of $a$ and $\theta$ for the
first excited state we also used the same procedure as in Gwa and Spohn
\cite{spohn} in order to show analitically that $z = \frac{3}{2}$,
$a_0= 2.301\;345\;96 \ldots$, independently of the value of $\tilde s$ and
$b_0 = \pi(\tilde s -1)/(2(\tilde s + 1))$. In the last line of Table 1 we
show the exact results obtained analitically.
These results indicate that all these models with arbitrary mixture if
molecules of different sizes, as well as the corresponding generalized
growth models exhibit a universal behavior with a KPZ-type of dynamical
behavior. In the appendix we show that for general values of $\epsilon_+,
\epsilon_-$ and $n$ the wave functions of \rf{6} for arbitrary distributions
of molecule's sizes, are exactly related. This implies that conditional
probabilities and correlation functions for arbitrary distribution $\{s\}$
are exactly related to the simple exclusion problem $\{s_1 = s_2 = \cdots
s_n =1\}$. The eigenvalues of these models are exactly related in the case
of free boundaries. In the case we have a periodic lattice the eigenvalues
of $H^{\{s\}}$ are exactly related to the asymmetric XXZ chain with
twisted boundary condition $\phi$ proportional to the momentum of the
first excited state. Since the momentum of this state is $P =2\pi/N$ the
effect of the twisted angle should not affect the leading order in the mass
gap calculations. This implies that for arbitrary values of $\epsilon_+,
\epsilon_-$ and density $n$ the leading order results of the real part
of the gap are the same as those calculated systematically in \cite{kim}.
\section{Concluding remarks}
We have solved exactly a general asymmetric diffusion problem where the
particles may have distinct and arbitrary integer sizes.
We also show in Sec. 2 that these diffusion
models are related to generalized growth models. Since through diffusion
the particles do not interchange positions, a
given order $\{s_1,\ldots,s_n\}$ of the distribution of molecule's sizes on
the lattice is fixed, up to cyclic permutations. A parameter which is
proportional to the excluded volume for the particles is
the average size of the molecules $\tilde s$ given
by \rf{54}. In the case of the simple exclusion problem all the molecules
have the same unity size $s_1 = s_2 =\cdots =s_n = 1$, which gives
$\tilde s = 1$. On the other hand, if all the particles has no size we have
$\tilde s =0$ and there is no excluded volume. By choosing suitable
distribution $\{s\}$ of molecule's sizes we may change $\tilde s$ almost
continuously in the bulk limit $N \rightarrow \infty$. Exploiting the
connection between the dyanmical critical exponent $z$ and the mass gap
of the related quantum chain we obtained $z = \frac{3}{2}$ for all the
models, independently of the parameter $\tilde s$ measuring the excluded
volume. This implies that, at least in one dimension, the excluded
volume effect is irrelevant for dynamical systems in the KPZ universality class.
We also show (see the appendix) that the wave functions of the models
with arbitrary distribution of molecule's sizes can be related to those of a
simple asymmetric exclusion problem, in a distinct lattice size. This
implies that conditional probabilities and correlation functions of these
models are exactly related.
\acknowledgements {
This work was supported in part by Conselho Nacional de Desenvolvimento
Cient\'{\i}fico e Tecnol\'ogico - CNPq - Brasil, by FINEP - Brasil and
by the Russian Foundation of Fundamental Investigation ( Grant 99-02-17646). }
|
\section{Introduction}
\label{sec:intro}
The recent finding that the electroweak phase transition might be only
a smooth cross\-over~\cite{Kajantie:1997qd} has changed the picture of
cosmological phase transitions in a fundamental way and shown that theories
with gauge fields can behave very differently from those with only
global symmetries. However, the statement about the smoothness of the
transition refers only to the equilibrium properties, and from the point
of view of cosmology, it is very important to understand also the
time evolution of the transition, in particular, whether the fields
fall out of equilibrium and what its consequences would be. For this,
a calculation scheme is needed for dynamical quantities that
treats the gauge fields correctly and gives the same results in
the special case of thermal equilibrium as the
the dimensional reduction method~\cite{Ginsparg:1980ef,Kajantie:1996dw}
that was used to show the
smoothness of the electroweak transition. In this paper, we propose
such a scheme.
Instead of the electroweak theory, we will concentrate here on
scalar electrodynamics, i.e.~the Abelian Higgs model.
Not only does the simpler gauge group allow more efficient simulations,
but the theory also contains topological defects, namely
Nielsen-Olesen vortices~\cite{Nielsen:1973ve},
and therefore it can be used to
simulate defect formation in a phase transition~\cite{Kibble:1976sj} and
to test the validity of the present predictions for the produced
defect density~\cite{Zurek:1996sj}
in gauge theories.
In Ref.~\cite{Vincent:1998cx},
the dynamics of the transition and, in particular, formation of vortices
was simulated classically on a lattice
with field equations derived from the original Lagrangian.
A similar approach had been suggested earlier~\cite{Grigoriev:1988bd}
in the context of electroweak baryogenesis.
This was based on the observation that the quantum distribution
of the long-wavelength modes is essentially classical.
Since the time evolution of a classical
field theory is given by the equations of motion, this approach
allows non-perturbative numerical simulations. For baryogenesis,
the interesting quantity is the hot sphaleron rate (see
Ref.~\cite{Rubakov:1996vz} and references therein), which is an
equilibrium quantity, but similar methods can also be used to
describe small deviations from equilibrium, such as phase transitions.
After all, the phase of the system is only a property of the long-wavelength
modes, and the
distribution of the high-momentum modes is to a large extent
independent of it. Therefore also the transition between the phases is
described by the classical long-wavelength modes.
This straightforward approach has the serious problem that
the results depend on the ultraviolet
cutoff\cite{Bodeker:1995pp,Arnold:1997yb}.
Because of the Rayleigh-Jeans ultraviolet divergences, a classical
continuum field theory cannot actually even be in thermal
equilibrium -- this was one of the reasons
why quantum mechanics was invented in the first place!
The divergences depend on the temperature and are absent at zero temperature,
so they cannot be
cancelled by introducing counterterms as in quantum theories.
Therefore it is not sufficient to
use the classical theory with the Lagrangian of the original
quantum theory, but one has to construct an effective Lagrangian for
the low-momentum (soft) modes by integrating out the high-momentum (hard)
modes. The effective Lagrangian then contains precisely the necessary
(temperature-dependent) counterterms to remove the divergences.
In calculating static properties, i.e.~expectation values of products
of operators measured
at the same time, this procedure is known as dimensional
reduction\cite{Ginsparg:1980ef}, and leads to a three-dimensional effective
theory, in which the temporal component of the gauge field
has a Debye mass term $m_D\sim gT$. To calculate non-static quantities, the
full four-dimensional effective Lagrangian needs to be constructed
by calculating the so-called hard thermal loops, and
in gauge theories it
turns out to be non-local~\cite{Silin:1960,Pisarski:1989vd}.
This makes
computer simulations practically impossible, since one would have to
keep in the memory all the configurations encountered during the time
evolution.
Fortunately, the equations of motion can be written in a local form
by introducing auxiliary fields\cite{Nair:1994xs,Blaizot:1993zk}.
The Abelian version of the resulting system of differential equations is
known in plasma physics as the Vlasov equation and describes collisionless
electron plasma. Numerical solution of the equations of motion is
still difficult, since the field $W$ depends not only on the coordinate
but also on the momenta of the hard particles it describes, and
therefore the theory becomes essentially five-dimensional.
The traditional method is to approximate it by a
large number of point particles.
This approach has been used to determine the hot sphaleron rate in the
electroweak theory in Ref.~\cite{Moore:1998sn}.
In this paper, we show how the theory can be formulated in such a way
that simulations in terms of fields rather than particles becomes
feasible. One of the problems we have to face is the multiplication of
the number of degrees of freedom when the hard modes are included. We show
that the number of extra degrees of freedom per lattice site must be chosen
carefully if the simulation of a particular soft mode is to be trusted:
if one wants to reproduce Landau damping in a mode of momentum
$k$, naive methods require of order $(kt)^2$ degrees of freedom per lattice site for the
hard modes. However, we are able to rewrite the
effective hard thermal loop improved
classical equations of motion in such a way that we only need of order
$kt$ extra degrees of freedom per lattice site. Our numerical methods
reproduce known analytic results in pure Abelian gauge theory extremely well, which
gives us confidence that we will be able to tackle the Abelian Higgs model
accurately.
The structure of the paper is the following. In Sec.~\ref{sec:kinetic} we
review the hard thermal loop Lagrangian and derive the Hamiltonian
formulation for the hard modes. In Sec.~\ref{sec:simu} we explain how
to approximate the system with a finite number of degrees of freedom
in numerical simulations. In Sec.~\ref{sec:anaresults} we derive some
analytical results, which we compare to our numerical results in
Sec.~\ref{sec:numerical}. Extending our technique to include the soft
Higgs field is discussed in Sec.~\ref{sec:higgs}, and Sec.~\ref{sec:disc}
contains the conclusions. The paper also has one appendix, in which
the continuum and lattice equations of motion for the Legendre modes
are given explicitly.
\section{Kinetic formulation}
\label{sec:kinetic}
Let us consider scalar electrodynamics at a temperature $T$. The Lagrangian
of the theory is
\begin{equation}
{\cal L}=-\frac{1}{4}F_{\mu\nu}F^{\mu\nu}+|D_\mu\phi|^2
-m^2|\phi|^2-\lambda|\phi|^4,
\labx{equ:lagr}
\end{equation}
where $D_\mu=\partial_\mu+ieA_\mu$.
We assume that $e\ll 1$ and $\lambda\sim e^2$ and that $m\ll T$ so that
high-temperature approximation can be used.
The phase structure of the theory
was determined in
Refs.~\cite{Dimopoulos:1997cz,Kajantie:1997hn}. When
the Higgs self-coupling
is small, there
is a first-order phase transition, but if
the self-coupling
is large enough,
the transition becomes continuous. Unlike in
the electroweak theory, it does not become a smooth
crossover. In Ref.~\cite{Kajantie:1998zn},
it was pointed out that this can be interpreted as a consequence
of the existence of Nielsen-Olesen vortices.
To one-loop order, the only non-vanishing contributions to the effective
Lagrangian arise from the two-point diagrams.
The correct procedure
would be to calculate the diagrams in
the original quantum theory and to write down a classical Lagrangian
that gives the same result when the diagrams are calculated on
a lattice~\cite{Bodeker:1995pp}. However, this is
not possible in practice, since the form of the necessary
effective lattice
Lagrangian is not known and is presumably very complicated because
the lattice has much less symmetry than the continuum.
In sphaleron rate calculations,
the effective Lagrangian
was taken from the continuum theory and the parameters were fixed by
matching only the static quantities~\cite{Moore:1998sn}. For the simulations
presented in this paper, this problem does not arise, as will be discussed
later.
In the high-temperature
approximation, the one-loop scalar self-energy
is simply
\begin{equation}
\Sigma=-\left( \frac{e^2}{4}+\frac{\lambda}{3}\right)T^2.
\end{equation}
The photon self-energy~\cite{Kraemmer:1995az}
is more complicated and if the external momentum is
$K=(\omega,\vec{k})$, it can be written in the form
\begin{equation}
\Pi^{\mu\nu}=\Pi_TP_T^{\mu\nu}+\Pi_LP_L^{\mu\nu},
\end{equation}
where $P_T^{\mu\nu}$ and $P_L^{\mu\nu}$ are the spatially
transverse and longitudinal
projection operators
\begin{equation}
P_T^{ij}=
\frac{k^ik^j}{k^2}-\delta^{ij},\quad P_T^{\mu 0}=0,\quad
P_L^{\mu\nu}=g^{\mu\nu}-\frac{K^\mu K^\nu}{K^2}-P_T^{\mu\nu},
\end{equation}
and
\begin{equation}
\Pi_L=m_D^2\left(1-\frac{\omega^2}{k^2}\right)
\left(1-\frac{\omega}{2k}\ln\frac{\omega+k}{\omega-k}\right),\quad
\Pi_T=\frac{1}{2}\left(m_D^2-\Pi_L\right).
\end{equation}
The Debye mass has the value $m_D^2=\frac{1}{3}e^2T^2+\delta m_D^2$, where
$\delta m_D^2$ is a cutoff-dependent counterterm
The self-energies can be resummed to a simple effective
Lagrangian for the soft modes
\begin{eqnarray}
{\cal L}_{\text{HTL}}&=&-\frac{1}{4}F_{\mu\nu}F^{\mu\nu}
-\frac{1}{4}m_D^2
\int\frac{d\Omega}{4\pi}F^{\mu\alpha}
\frac{v_\alpha v^\beta}{(v\cdot\partial)^2} F_{\mu\beta}\nonumber\\
&&+|D_\mu\phi|^2
-m_T^2|\phi|^2-\lambda|\phi|^4,
\labx{equ:lagrHTL}
\end{eqnarray}
where $m_T^2=m^2+(e^2/4+\lambda/3)T^2+\delta m_T^2$, and the integration
is taken over the unit
sphere of velocities $v=(1,\vec{v})$, $\vec{v}^2=1$.
Note that the form of $m_T^2$ justifies using the high-temperature
approximation even slightly below
the transition. At tree level, the transition
takes place when $m_T^2=0$, which shows that $m^2\sim -e^2T_c^2$.
The minimum of the potential is therefore at $v\sim T_c$, and the mass given
by the Higgs mechanism is $m_H^2\sim e^2T_c^2\ll T^2$.
The equations of motion derived from the Lagrangian (\ref{equ:lagrHTL})
are
\begin{mathletters}
\begin{eqnarray}
\partial_\mu F^{\mu\nu}&=&m_D^2\int\frac{d\Omega}{4\pi}
\frac{v^\nu v^i}{v\cdot\partial}E^i-2e{\rm Im}
\phi^*D^\nu\phi,
\labx{equ:nonlA}
\\
D_\mu D^\mu\phi&=&-m_T^2\phi-2\lambda(\phi^*\phi)\phi.
\labx{equ:nonlf}
\end{eqnarray}
\end{mathletters}
The derivative in the denominator in Eq.~(\ref{equ:nonlA}) means that
in order to simulate it numerically, one needs to keep the whole time
evolution of the system in the memory at the same time, which is impossible.
The form of the
equation of motion
(\ref{equ:nonlf}) of the scalar field is much simpler, and actually the only
effect of the hard modes is the modified mass term.
The scalar field can therefore be simulated with standard methods and
we will neglect it from now on, concentrating only on the gauge field.
We will refer to this theory as the pure gauge theory although
it contains the contribution from the hard scalar modes.
The non-locality problem of Eq.~(\ref{equ:nonlA})
can be solved by introducing new
fields\cite{Nair:1994xs,Blaizot:1993zk}. We follow Ref.\cite{Blaizot:1993zk}
and add the field $W(x,\vec{v})$,
which satisfies the equation of motion
\begin{equation}
(v\cdot\partial)W(x,\vec{v})=\vec{v}\cdot\vec{E}(x),
\labx{equ:eomW}
\end{equation}
and replacing Eq.~(\ref{equ:nonlA}) with
\begin{equation}
\partial_\mu F^{\mu\nu}=j_W^\nu(x)+j^\nu(x),
\labx{equ:eomA}
\end{equation}
where
\begin{equation}
j_W^\nu(x)=m_D^2\int\frac{d\Omega}{4\pi}v^\nu W(x,\vec{v}),
\labx{equ:currW}
\end{equation}
and $j^\nu(x)$ denotes the current due to the scalar field and any
external currents.
The system of equations (\ref{equ:eomW}), (\ref{equ:eomA}) is
a special case of
what is known as the
Vlasov equation in plasma physics, where it describes collisionless
electron plasma. The field $W(x,\vec{v})$ gives
the deviation of the density of
hard particles of velocity $\vec{v}$ from their equilibrium distribution.
In addition to the equations of motion, we also have to specify the
initial conditions, and since we would like to have the system initially
in a thermal equilibrium, they should be given by the equilibrium
distribution, i.e.~we should have a large number of initial configurations
with a Boltzmann distribution
$\propto\exp(-\beta H)$, where $H$ is the Hamiltonian. However, the equation
of motion (\ref{equ:eomW}) for the $W$ field is not of canonical form,
which makes it difficult to find the correct Hamiltonian.
The suggestion of\cite{Iancu:1997md} was to use the
conserved quantity
\begin{equation}
H=\frac{1}{2}\int d^3x\left\{\vec{E}^2+\vec{B}^2+
m_D^2\int\frac{d\Omega}{4\pi}W(x,\vec{v})^2
\right\}.
\labx{equ:hamW}
\end{equation}
In the simulations, one has to approximate the system with a finite
number of degrees of freedom. Normally, this is done by discretizing the
space to a finite lattice, but in our case, the field $W$ also depends on a
two-dimensional continuous parameter $\vec{v}$. There are various ways to
handle it in simulations. In Refs.~\cite{Moore:1998sn,Hu:1997sf}, $W$ was
approximated with a large number of particles. The other straightforward
options are approximating the velocity integral with a discrete set
of velocities on the unit sphere, or expanding $W$ in terms of
spherical harmonics. However, as we will show next, at least in the Abelian
case it pays to simplify the problem a bit first.
Let us define in Fourier space
\widetext
\begin{eqnarray}
\vec{f}(\omega,\vec{k},z)&=&
im_D\sqrt{\frac{2z^2}{1-z^2}}\int\frac{d\Omega}{4\pi}
\frac{\vec{v}\times\vec{k}}{(\vec{v}\cdot\vec{k})^2}
W(\omega,\vec{k},\vec{v})\left(
\delta(\vec{v}\cdot\hat{k}-z)+\delta(\vec{v}\cdot\hat{k}+z)
\right)\nonumber\\&&
-\frac{im_D}{k^2}\sqrt{\frac{1-z^2}{2z^2}}\vec{k}\times\vec{A},\nonumber\\
\theta(\omega,\vec{k},z)&=&im_D\int\frac{d\Omega}{4\pi}
\frac{1}{\vec{v}\cdot\vec{k}}
W(\omega,\vec{k},\vec{v})\left(
\delta(\vec{v}\cdot\hat{k}-z)+\delta(\vec{v}\cdot\hat{k}+z)
\right)
+\frac{im_D}{k^2}\vec{k}\cdot\vec{A}.
\labx{equ:defft}
\end{eqnarray}
Then the equations of motion in the temporal gauge $A_0=0$ are
\begin{mathletters}
\label{equ:formulation}
\begin{eqnarray}
\partial_0^2{\vec{f}}(z)&=&z^2\vec{\nabla}^2\vec{f}+m_Dz\sqrt\frac{1-z^2}{2}
\vec\nabla\times\vec{A},\\
\partial_0^2{\theta}(z)&=&z^2\vec{\nabla}\cdot\left(
\vec{\nabla}\theta-m_D\vec{A}\right),\\
\partial_0^2{\vec{A}}&=&-\vec\nabla\times\vec\nabla\times\vec{A}
+m_D\int_0^1dzz^2\left(
\vec\nabla\theta-m_D\vec{A}+
\sqrt{\frac{1-z^2}{2z^2}}\vec\nabla\times\vec{f}
\right)+\vec{j}.
\end{eqnarray}
\end{mathletters}
One advantage of this formulation is that these equations of motion are
canonical and the corresponding Hamiltonian is
\begin{eqnarray}
H&=&\frac{1}{2}\int d^3x\int_0^1dz\Biggl[
\vec{E}^2+(\vec\nabla\times\vec{A})^2+\vec{F}^2+{\Pi}^2+
z^2(\vec\nabla\times\vec{f})^2+z^2(\vec{\nabla}\theta-m_D\vec{A})^2
\nonumber\\&&
-m_Dz\sqrt\frac{1-z^2}{2}\vec{f}\cdot\vec\nabla\times\vec{A}
\Biggr],
\labx{equ:hamf}
\end{eqnarray}
\narrowtext
where $\vec{F}=\partial_0{\vec{f}}$ and $\Pi=\partial_0\theta$
are the canonical momenta of $\vec{f}$ and $\theta$,
respectively. We also need two extra conditions, namely the transverseness of
$\vec{f}$ and Gauss's law
\begin{eqnarray}
\vec\nabla\cdot\vec{f}&=&\vec\nabla\cdot\vec{F}=0,\nonumber\\
\vec\nabla\cdot\vec{E}&=&-m_D\int_0^1dz\Pi(z).
\labx{equ:Gauss}
\end{eqnarray}
By reformulating the degrees of freedom, we have gained two important
advantages. First, $\vec{f}$ depends only on one internal coordinate whereas
$W$ depends on two, and second, since we know the Hamiltonian
(\ref{equ:hamf}), we know the correct distribution of the initial
configurations.
\section{Simulations}
\label{sec:simu}
We still have to approximate the $z$-dependence of the hard modes with a
finite number of degrees of freedom, and we will use Legendre polynomials
for that.
We define
\begin{eqnarray}
\vec{f}^{(n)}&=&\int_0^1dzz\sqrt\frac{1-z^2}{2}P_{2n}(z)\vec{f}(z),
\nonumber\\
\theta^{(n)}&=&\int_0^1dzz^2P_{2n}(z)\theta(z),
\labx{equ:defLeg}
\end{eqnarray}
where $P_{n}(z)$ is the $n$th Legendre polynomial. The equations
of motion for different Legendre modes are given in Eq.~(\ref{equ:eomLeg})
in the appendix. It is also shown in the appendix that the approximation
can be trusted if the simulation time is less than
\begin{equation}
t_0\approx 2N_{\text{max}}/k.
\labx{equ:reltime}
\end{equation}
On the lattice, the field $\theta$ is defined on lattice sites, while
$\vec{f}$ and the gauge field $\vec{A}$ are defined on links between
the sites in
such a way that the invariance under time-independent gauge
transformations is preserved. Note that the field $\theta^{(0)}$ is
not gauge invariant by itself, but the combination
$\vec\nabla\theta^{(0)}-\frac{1}{3}m_D\vec{A}$ is.
The canonical momenta are defined at half-way between the time steps so
that the value of, say, $\vec{f}$ at time $t+\delta t$ is determined
from the fields at time $t$ and the momenta at time $t+\delta t/2$.
In this way, the time reversal invariance of the continuum theory is
preserved. The lattice equations of motion are given in
Eq.~(\ref{equ:eomLatt}) in the appendix.
The aim of the simulations described here was mainly to
test the simulation code by comparing it with known analytical results.
Therefore we neither averaged over thermal initial conditions nor
included the scalar field.
The equations of motion for the pure gauge theory
are linear and can therefore in principle be solved analytically.
It also means that the modes with different momenta do not interact
with each other, and to simulate a mode with momentum along, say, the
$z$ axis, we can use a lattice with $1\times 1\times N_z$ sites. This
makes the simulations very simple. Moreover, the pure gauge theory is
ultraviolet finite, so no mass counterterms are needed.
In particular, $\delta m_D^2$ and $\delta m_T^2$ introduced in
Sec.~\ref{sec:kinetic} vanish.
\section{Analytical Results}
\label{sec:anaresults}
Before specializing to particular solutions, we first discuss the
finite temperature propagator for the pure Abelian theory.
Due to linearity, it is sufficient to consider a single
Fourier mode at a time. Let the momentum be $\vec{k}$, and, for simplicity,
let us consider just the transverse part of the propagator $G(t,\vec{x})$.
Its Fourier transform
has a non-analytic form
\begin{equation}
\frac{1}{G(\omega,\vec{k})}=
-\omega^2+k^2+\frac{m_D^2}{4}\left[
2\frac{\omega^2}{k^2}+\frac{\omega}{k}\frac{k^2-\omega^2}{k^2}
\ln\frac{\omega+k}{\omega-k}
\right].
\labx{equ:prop}
\end{equation}
The analytic structure of the propagator is shown in Fig.~\ref{fig:prop}.
The propagator has two oscillatory poles on the real axis at
$\omega^2_p\approx k^2+\frac{1}{3}m_D^2$, but also a branch cut from
$\omega=-k$ to $\omega=k$.
The original equation of motion (\ref{equ:nonlA}) implies that the branch
cut must be taken along the real axis. There are no other singularities
on this physical Riemann sheet, but if one continues the propagator
analytically from the upper half-plane through the branch cut,
one finds a pole at
\begin{equation}
\omega=-i\gamma_L \approx -i\frac{4k^3}{\pi m_D^2}.
\end{equation}
Likewise, continuing analytically from the lower half-plane through the
cut reveals a pole at $\omega=i\gamma_L$.
The solution to the inhomogeneous equation with an external
source
$\vec{j}(\omega)$ is, assuming that the current is transverse
and that the fields vanish at $t=-\infty$,
\begin{equation}
\vec{A}(t,\vec{k})=
\int_{-\infty+i\epsilon}^{\infty+i\epsilon}\frac{d\omega}{2\pi}
e^{-i\omega t} G(\omega)\vec{j}(\omega).
\labx{equ:formalsol}
\end{equation}
If $t>0$, the integral can be transformed into a contour integral by
closing it on the lower half-plane.
This integral has three pieces: two contributions from
the poles and one from the integral around the branch cut.
The result of the integration around
the cut is sensitive to the functional form of $\vec{j}(\omega)$, a
reflection of the non-locality of Eq.~(\ref{equ:nonlA}): the behavior of
the system depends on its whole history.
For example, Boyanovsky et al.~\cite{Boyanovsky:1998pg} discussed a situation
in which an inhomogeneous initial configuration is set up for the
soft fields $\vec{A}$ and $\vec{E}$, and calculated its relaxation
to the equilibrium at asymptotically
long times. Still keeping the assumption of transverseness,
the solution was given as a function
of $\vec{A}$ and $\vec{E}$ at time $t=0$,
\begin{eqnarray}
\vec{A}(t,\vec{k})&\approx&
\vec{A}(0,\vec{k})
\left[Z[T]\cos\omega_pt-\frac{4}{m_D^2}\frac{\cos kt}{t^2}\right
-\vec{E}(0,\vec{k})
\left[Z[T]\frac{\sin\omega_pt}{\omega_p}-
\frac{4}{m_D^2}\frac{\sin kt}{kt^2}\right],
\labx{equ:boyaresult}
\end{eqnarray}
where
\begin{equation}
Z[T]=-\left(\frac{\partial G^{-1}}
{\partial \omega^2}\right)^{-1}_{\omega=\omega_p}.
\end{equation}
Eq.
(\ref{equ:boyaresult}) can be seen to correspond to
$\vec{j}(\omega)=-i\omega\vec{A}(0)-\vec{E}(0).$
As they emphasize, the dominant contribution does not come from
the smallest frequencies but those near the end points of the branch cut.
With a different
choice of $\vec{j}$, qualitatively different solutions
can be found. Suppose that $\vec{j}(t)$ is increased very
slowly from zero to a finite value and then suddenly switched off,
i.e.~$\vec{j}(t)=\vec{j}_0e^{\gamma_0 t}\Theta(-t)$,
where $\gamma_0$ is eventually
taken to zero and $\Theta(t)$ is the step function.
The Fourier transform $\vec{j}(\omega)=\vec{j}_0/(\gamma_0+i\omega)$
is peaked around the origin, and therefore the integral around the branch
cut is dominated by the integrand near $\omega=0$. Hence, expanding
the propagator around $\omega=i\epsilon$ and $\omega=-i\epsilon$ on
the upper and lower half-planes, respectively, we may write
\begin{equation}
\left.\vec{A}(t,\vec{k})\right|_{\rm cut}\approx
\vec{j}_0\int_{-k}^k\frac{d\omega}{2\pi}\left ( e^{-i\omega t}
\frac{1}{\gamma_0+i\omega}\frac{1}{\gamma_L - i\omega} +
e^{i\omega t}
\frac{1}{\gamma_0-i\omega}\frac{1}{\gamma_L - i\omega}
\right)
\frac{\gamma_L}{k^2}.
\end{equation}
The
limits of the integral may be taken to infinity, and the contour closed
in either the upper or the lower half-plane depending on the sign of the
exponent. The second term has no poles in the upper half-plane and hence
vanishes, while the first term gets a contribution from the pole at
$\omega = -i\gamma_L$.
Thus we find an exponentially damping solution
\begin{equation}
\left.\vec{A}(t,\vec{k})\right|_{\rm cut}
\approx -\frac{\vec{j}_0}{k^2} e^{-\gamma_Lt}.
\end{equation}
This exponential decay is known as Landau damping, and it is important to
notice that it did not really come from integrating around a
pole of the propagator
(\ref{equ:prop}). If it had, then the equations of motion would
either have to break time reversal invariance, which they do not do,
or there would have to be a corresponding exponentially growing
solution, thus making the system unstable.
For completeness, we also calculate the contribution coming from the
poles at $\omega=\pm\omega_p$. The full result is then
\begin{equation}
\vec{A}(t,\vec{k})\approx
-\vec{j}_0\left(
Z[T]
\frac{\cos\omega_pt}
{\omega_p^2}
+
\frac{1}{k^2} e^{-\gamma_Lt}
\right).
\labx{equ:landauasymp}
\end{equation}
\section{Numerical results}
\label{sec:numerical}
In order to numerically
reproduce the results (\ref{equ:boyaresult}), (\ref{equ:landauasymp}),
we have to specify
the correct initial conditions in terms of $\vec{f}$ and $\theta$.
The result (\ref{equ:boyaresult}) is given simply by $\vec{f}(0)=\theta(0)=0$.
For the soft modes, we chose
$\vec{A}(0)=0$, $\vec{E}(0)=10\hat{y}\sin \vec{k}\cdot\vec{x}$.
To see the power-law damping most
clearly, $m_D$ must be relatively small, and we chose $m_D=2\pi$,
$\vec{k}=2\pi\hat{x}$ in
our units. The lattice size was $20\times 1\times 1$, lattice spacing
$a=0.05$ and the time step $\delta t=0.01$. We used $N_{\rm max}=200$
so that according to Eq.~(\ref{equ:reltime})
the results should be reliable when $t\lesssim 60$.
Eq.~(\ref{equ:boyaresult}) becomes
\begin{equation}
\vec{A}(t,\vec{k})\approx
\left(-1.236\sin 7.451x+0.1613\frac{\sin 2\pi t}{t^2}\right)\hat{y}.
\labx{equ:boyanum}
\end{equation}
We added to the numerical result a constant-amplitude part
$\alpha\sin\beta t$ and determined the parameters $\alpha$ and
$\beta$ from the condition that the remaining part agrees with the
decaying part of Eq.~(\ref{equ:boyanum}). The result with $\alpha=1.24008$
and $\beta=7.43024$ is shown in Fig.~\ref{fig:boyacmp}. The discrepancy
between the numerical and analytical results is less than one percent.
The initial conditions that correspond to Eq.~(\ref{equ:landauasymp})
are those in which the first and second time derivatives of the fields
vanish. If we choose $\vec{j}_0$ such that
$\vec{A}(0,\vec{x})=\hat{y}\sin \vec{k}\cdot\vec{x}$,
then Eqs.~(\ref{equ:defft}) and
(\ref{equ:defLeg})
imply that
\begin{eqnarray}
\vec{f}^{(0)}(0,\vec{x})&=&\frac{m_D}{3k}\hat{z}
\cos \vec{k}\cdot\vec{x},\nonumber\\
\vec{f}^{(1)}(0,\vec{x})&=&-\frac{m_D}{15k}\hat{z}
\cos\vec{k}\cdot\vec{x},\nonumber\\
\vec{f}^{(n>1)}(0,\vec{x})&=&0,\quad\theta^{(n)}(0,\vec{x})=0.
\end{eqnarray}
The exponential damping is seen most clearly if $m_D\gg k$, and we chose
two different values $m_D=10\pi$ and $m_D=20\pi$ while we took again
$\vec{k}=2\pi\hat{x}$.
Again, the lattice size was $20\times 1\times 1$, the lattice spacing
$a=0.05$, the time step $\delta t=0.01$, and $N_{\rm max}=200$. The results
of the simulations are shown in Fig.~\ref{fig:landaucmp}. The predicted decay
rates are $\gamma_L=0.32$ and $\gamma_L=0.08$ and the amplitudes of the
oscillation $A_{\rm osc}\approx 0.103$ and $A_{\rm osc}\approx0.029$.
As can be seen from Fig.~\ref{fig:landaucmp}, these agree very well with the
numerical results.
We also carried out simulations with with different values of $N_{\rm max}$
to find out at what time the approximation breaks down and to
test the estimate (\ref{equ:reltime}). We used $m_D=20\pi$ and the values
for the other parameters were as before.
Since all the Legendre modes with $n>1$ are
initially zero, we expect the first errors to occur at time
$t\approx 2t_0\approx 4N_{\rm max}/k$.
The result is shown in Fig.~\ref{fig:testNmax}
for various values of $N_{\rm max}$ and confirms our estimate.
The number of extra scalar degrees of freedom needed for simulating
with any particular value of
$N_{\rm max}$ is $8N_{\rm max}$.
In order to compare the efficiency of the Legendre polynomial formulation
with other approaches, we carried out
the same simulations with a more straightforward method. We chose a
large number of points on the unit sphere to represent different velocities
and used them to simulate the pair of equations (\ref{equ:eomW}),
(\ref{equ:eomA}). More precisely, the different velocities were
\begin{equation}
\vec{v}=\frac{(N_v+\frac{1}{2},
n_y,n_z)}{\sqrt{(N_v+\frac{1}{2})^2+n_y^2+n_z^2}},
\quad -N_v\le n_i\le N_v,
\end{equation}
and those obtained from that with reflections or rotations of $\pi/2$.
The parameters were the same as in Fig.~\ref{fig:testNmax}.
The values of $N_v$ used ranged from 2 to 8, and the corresponding
number of extra degrees of freedom is $6(2N_v+1)^2$.
The results in Fig.~\ref{fig:testvkoko} show clearly
that the number of different velocities becomes quickly prohibitive
when the simulation time is increased.
If we assume that, as with the Legendre polynomials, the particularly
smooth initial conditions result in a factor of two in the reliable
simulation time, we can
estimate that generally
a simulation will be reliable if $t\lesssim \pi N_v/k$.
This estimate can also be reached analytically. Since the time evolution is
a superposition of oscillations, the reliable simulation time is
$t_0\approx \pi/\omega_0$, where $\omega_0$ is the smallest frequency,
i.e.~the pole of the propagator that is nearest to the origin.
The transverse self-energy $\Pi_T(\omega)$ diverges if
$\omega=\vec{k}\cdot\vec{v}$ for any velocity
$\vec{v}$.
The smallest pole of the propagator is at a frequency $\omega_0$, which is
smaller than $\min\{\vec{k}\cdot\vec{v}\}$, but of the same
order of magnitude. Generally $\min\{\vec{k}\cdot\vec{v}\}
\approx k/N_v$, leading to the previous estimate.
Since expansion of $W$ in
Eq.~(\ref{equ:eomW}) in terms of spherical harmonics $Y^m_n$
is essentially equivalent to the expansion of $\vec{f}$ and $\theta$
in terms of Legendre polynomials, we can also estimate the effectiveness
of that approach. In order to reach the same simulation time,
the highest $n$ used should be $2N_{\rm max}$.
The number of extra degrees of freedom would then be
$(2N_{\rm max}+1)^2$, which again increases much faster than in
the Legendre polynomial approach.
Our estimates of the reliable simulation times
in the different approaches have been summarized in table~\ref{tab:simtimes}.
They show that the Legendre polynomial formulation is the most
economical one.
\footnote{
The analogous number of degrees of freedom in the particle
method used by Moore et al.~\cite{Moore:1998sn} is $6\langle n\rangle$.
The value of $\langle n\rangle$ they were using varied between 17 and 120,
but they did not carry out this kind of systematic analysis
of the corresponding reliable simulation time. However,
the intrinsic randomness of the method seems to imply that it does not
reproduce the correct
behavior as accurately at short times as the methods discussed here.}
\section{Higgs model}
\label{sec:higgs}
The pure gauge theory that has been discussed so far is an ideal
way to test the formalism, since it is linear and
one can in principle solve it exactly. In the previous section, we have
shown that the numerical results agree with the analytical calculations
and we were even able to estimate the number $N_{\rm max}$ of Legendre
modes needed to ensure that the simulation is reliable.
It is very easy to transform the system to a very non-trivial one simply
by adding a Higgs field. While this makes it impossible to solve
the equations of motion analytically, from the point of view of numerical
simulations it means only adding one more field. Still, the
resulting theory has a complicated structure with a phase transition and
topological defects.
As was shown in Sec.~\ref{sec:kinetic}, the weak-coupling condition
$e\ll 1$ and $\lambda\sim e^2$ implies that the hard thermal loop
approximation is also valid near the transition in the broken phase.
The reason is that the distribution of the hard modes is independent
of the phase.
Furthermore, when the phase transition takes place in
a finite time, only the soft modes fall out of equilibrium, and therefore
the classical description remains valid, even in such non-equilibrium
processes.
The above suggests that the classical formalism discussed in this paper
can be used to simulate non-equilibrium dynamics of the phase
transition in hot scalar electrodynamics.
In the cosmological setting the transition takes place when the universe
cools as it
expands, but in a simulation some other way to change the temperature
is needed. A straightforward way of doing this is preparing an initial
ensemble with a higher temperature for the soft modes than for the hard modes.
A large number of initial configurations would then be taken from this
ensemble and evolved in time. When the soft and hard modes interact,
the temperature of the soft modes decreases and they
undergo a phase transition.
While this kind of an instantaneous quench in the temperature is not
very realistic, it would still be a first approximation to the
true cosmological phase transition. These simulations would
answer many interesting questions about the dynamics of the phase
transition, for instance the density of topological defects created.
Certainly, the hard thermal loop
approximation we have used
is only valid at relatively
short times.
We have neglected the collisions of the hard
particles and at long times
$t\gtrsim 1/e^4T$
they would have an essential contribution
to the time evolution~\cite{Blaizot:1999xk}.
If the coupling is weak, as we assume, this is not a serious problem.
The other problem is to remove the effect of the ultraviolet lattice
modes as discussed in Sec.~\ref{sec:kinetic}.
The fact that the simulations
of \cite{Vincent:1998cx} showed no signs of a Rayleigh-Jeans catastrophe
developing over the course of the simulation
seems to indicate that the effect can safely be neglected.
Phase transitions and other non-equilibrium processes can be simulated
also in non-Abelian theories using the Vlasov equation (\ref{equ:eomW})
to describe the hard modes.
It would be important to know, whether a formulation analogous to
Eq.~(\ref{equ:formulation}) with only one internal
coordinate is also possible in non-Abelian theories, since
that would reduce the need of computational power drastically.
Since the derivative in Eq.~(\ref{equ:eomW})
is replaced with a covariant derivative,
operating in the momentum space becomes more complicated and
the same derivation cannot be used.
\section{Conclusions}
\label{sec:disc}
In this paper, we have studied hot Abelian gauge field theory in
the hard thermal loop approximation. Starting from the local kinetic
formulation (\ref{equ:eomW}), (\ref{equ:eomA}), we were able to
reformulate the degrees of freedom in a more economical way.
The equations of motion in the new formulation are
canonical and we found the explicit form of the Hamiltonian
(\ref{equ:hamf}).
The pure gauge theory discussed in this paper is linear and can therefore
be solved analytically, at least in principle. We pointed out that
because of non-locality it is not sufficient to specify the initial
conditions for the soft modes only. In a sense, one will have to specify
the whole history of the system in order to calculate its behavior in
the future. Taking this effect into account modifies the existing
results~\cite{Boyanovsky:1998pg} and leads to exponential damping
(\ref{equ:landauasymp})
with the Landau damping rate.
In order to simulate the system, we approximated the functions
$\vec{f}$, $\theta$ with a finite number of Legendre polynomials. The
simulations reproduced the analytical results, and the number of
degrees of freedom needed to describe the hard modes
was much smaller than in other possible approaches.
This shows that the
formulation presented in this paper is very well suited for numerical
simulations.
While the pure gauge theory is trivial in the sense that it can be solved
analytically, including the soft Higgs field makes analytical calculations
essentially impossible. On the other hand, it is very simple to add it to
numerical simulations. Leaving possible ultraviolet problems aside,
that would give a way to study non-perturbatively the
non-equilibrium dynamics of
a phase transition in a gauge field theory.
\acknowledgments
AR would like to thank D.~B\"odeker, D.~Boyanovsky, E.~Iancu, M.~Laine,
G.D.~Moore and
K.~Rummukainen for useful discussions. This work was supported by
PPARC grant GR/L56305.
AR was partly supported by the
University of Helsinki.
|
\section{Introduction}
The propagation of light from distant objects like galaxies and QSOs
is deflected by the gravitational tidal forces caused by inhomogeneous
matter distribution between the objects and us. The so-called lens
effect creates conspicuous images such as multiple QSOs and arcs in
clusters of galaxies, owing to special positional relations among sources,
lens objects and the observer, but in most cases it causes small
deformations and amplification of optical images, which result from
the superposition of deflections from many lens objects on
cosmological scales. This so-called weak lensing gives us valuable
information on the structure and evolution of the universe, especially
regarding the distribution of dark matter,
not only on large scales, but also on small scales around galaxies.
The statistical behavior of optical quantites such as
convergence, amplification and shear due to weak lensing
has been studied by many people since the pioneering papers
by Gunn,\cite{rf:gunna}\tocite{rf:gunnb} Weinberg\cite{rf:weina} and
Blandford and Jaroszy\'nski.\cite{rf:bj} For the derivations of these
quantities there have
been various approaches which consist of the multi-lens-plane method,
direct integration methods solving the null-geodesic equation,
optical scalar equations and
the equation of geodesic deviation, and perturbative methods.
\noindent 1.1. {\it Multi-lens-plane method}
This is one of the various methods numerically simulating
light propagation in
inhomogenous model universes, which were derived and developed by
Blandford and Narayan,\cite{rf:bn86}
Schneider and Weiss,\cite{rf:sw88a}\tocite{rf:sw88b} Jaroszy\'nski
et al.,\cite{rf:jppg} Jaroszy\'nski.\cite{rf:jaro91a}\tocite{rf:jaro92b}
In this method we first build a finite number of planes normal to light rays
between an observer and a source, and the three-dimensional
distribution of galaxies and dark matter as lens objects is replaced by
their two-dimensional distribution in the planes onto which the matter
in each interval is projected.
The deflection of light rays resulting from each lens plane is
calculated using geometrical optics, and the statistical averages of
optical quantities for many-ray bundles are derived. \cite{SCH92}
To use this method we must produce the inhomogeneous distribution of
lens objects in each lens plane. In early works (Schneider and
Weiss,\cite{rf:sw88b} Paczy\'nski and Wambsganss,\cite{rf:pw89} Lee and
Paczy\'nski.\cite{rf:lp90}) the random distribution of lens objects with
an equal mass was assumed, and so the realistic large-scale structure
of the matter distribution was neglected. Jaroszy\'nski et
al.\cite{rf:jppg} constructed an inhomogeneous model using an $N$-body
simulation (the PM code), but in this model large-scale structure was
considered, while small-scale structure corresponding to galaxies
was neglected.
Jaroszy\'nski\cite{rf:jaro91a}\tocite{rf:jaro92b} developed other numerical
models of matter distribution using the Zeldovich approximation,
located galaxies (as lens objects) correponding to the matter density,
and assumed for them the mass spectrum due to the Schechter luminosity
function and their morphological types randomly.
Wambganss et al.\cite{rf:wam95} \tocite{rf:wam98} generated models for
the evolution of large-scale structures by $N$-body simulation in
the CDM models, investigated their
influence on light propagation and statistical results such as
the frequency and separation angles of multiple QSOs and the image
deformation of high-redshift objects. In their works, galaxies as lens
objects were not taken into account.
Recently Premadi et al.\cite{rf:prem98} have improved the work of
Jaroszy\'nski et
al. by adopting an $N$-body simulation (P$^3$M code) and
considering the spatial distribution of galaxies with the mass
spectrum and morphological types (due to the morphological
type-density relation). They adopted Jaroszy\'nski's assumption that the
main lens objects are galaxies and the background
matter outside galaxies has smooth distribution with radii
$\approx 1 h^{-1}$ Mpc. In \S 2 the recent result of
studies by Premadi et al.\cite{rf:prem98} is given.
\noindent 1.2. {\it Direct integration methods}
There are three types of direct integration methods in which the
optical scalar equation (derived by Sachs\cite{rf:sachs61}), the
equation of geodesic deviation and the null-geodesic equation are
solved. Kantowski's approach belongs to the first type. He derived
the averaged optical scalar equation and solved it in an inhomogeneous
model universe with swiss cheese structure.\cite{rf:kant69}
Dyer and Roeder\cite{rf:dr3} studied the behavior of the
amplification applying this approach to the case of a nonzero
cosmological constant. Recently, Kantowski et al.\cite{rf:kant95} have
analyzed the lens effect in the observed redshift-magnitude relation
for type Ia supernovae.
Watanabe and Tomita\cite{rf:wt90} studied the behavior of shear in
the ray bundles in various models in which the random distribution of
particles simulating galaxies and clusters of galaxies was assumed
for simplicity. As an example of the second type of direct integration
method (treating the equation of geodesic deviation)
we cite a recent work of Holz and Wald,\cite{rf:hw98} who studied
the behavior of amplification in an inhomogeneous model with a
swiss cheese-like structure. In their case not all systems consisting of a
central galaxy and the surrounding empty region have the
compensated mass distribution, in contrast to Kantowski's swiss cheese
system. They derived the probability distribution of amplification,
which was found to be non-Gaussian in accord with
Dyer-Oattes\cite{rf:dyoa83} and examined the significance of the amplification
on the observation of type Ia supernovae (Holz\cite{rf:holz98}).
A drawback of this approach may be that it treats only ray bundles with
infinitesimal separation angles.
The third type of direct integration method is described in the works
of Tomita and Watanabe.\cite{rf:tw89}\tocite{rf:tw90} They treated the
multiple deflection of light rays directly by solving the
null-geodesic equation and studied the statistical averages of image
deformation in inhomogeneous universe models. The evolution of models
was derived under the periodic condition using Aarseth's individual
particle method for $N$-body simulations, in which particles are
regarded as galaxies or clusters of galaxies with the initial power
spectrum $n = 0$.
The particles (with the same mass and particle number $N = 1331$)
are placed and move in a periodic box with comoving length $\approx
50$ Mpc, and light rays also propagate in the box under
gravitational forces from particles in the box.
In Tomita and Watanabe,\cite{rf:tw89} the influence of lensing on the
CMB anisotropy was analysed and a negative result was obtained,
contrary to expectations (Kashlinsky,\cite{rf:ka88}
Tomita,\cite{rf:tom88}\tocite{rf:tpr89} Sasaki.\cite{rf:sasaki89})
In the latter paper,\cite{rf:tw90} examples of strong image
deformations and their statistics were studied.
Using the same approach, Tomita\cite{rf:tom98a}\tocite{rf:tom98c} has
recently studied the statistical behavior of optical scalars in
more realistic inhomogeneous models with the CDM spectrum ($n = 1$).
All particles are assumed to have the same galactic mass, while their
radii depend on the model used.
In the first two of these papers,\cite{rf:tom98a}\tocite{rf:tom98b}
the lens effect was
analyzed in flat and open models, respectively, and in the third
paper,\cite{rf:tom98c} the distribution of angular diameter distances
in these models was derived.
As for the number density of main lens
objects, there is an important difference between the work of Premadi
and that of Tomita. In the former, adopting Jaroszy\'nski's assumption
mentioned above, the density of main lens objects (galaxies) is found
to be $\Omega_g
\approx 0.02$. In Tomita's works, on the other hand, the density
($\Omega_L$) of main lens objects (called {\it compact lens
objects}) is $\Omega_L = 0.2$ for three models with $\Omega_0 = 1$
and $0.2$
and $\Omega_L = 0.4$ for a model with $\Omega_0 = 0.4$, which are much larger
than $\Omega_g$. This difference between these two works comes from
the different estimation for the role of non-galactic clouds as lenses.
At present it is not clear whether Jaroszy\'nski's assumption is
realistic or
not, and so the matter outside galaxies may not be so smooth and may
behave as discrete clouds with radii $r_{\rm cl} \approx 100$ kpc,
consisting of dark matter and the baryon fluid. In these clouds a
baryon fluid has experienced no dissipated collapse and has less
central concentration than in galaxies, but as weak lens objects they
may play a role similar to galaxies.
An outline of Tomita's recent works is given in \S 3, and the results of
additional analyses are also given in the case (satisfying Jaroszy\'nski's
assumption) that the density of lens
objects is $\Omega_L = \Omega_g$ and the remaining objects are
assumed to be smoother, by constraining their radii so as to be $500
h^{-1}$kpc. By comparing the two results in the case of $\Omega_L = 0.2$
and the case of $\Omega_L = \Omega_g$, we find how lensing on the
small scale is
dominated by non-galactic lens objects, if they exist.
\noindent 1.3. {\it Perturbative methods}
The method for solving optical scalar equations perturbatively with respect
to a small perturbed expansion and shear has been introduced by
Gunn\cite{rf:gunnb} and developed by Babul and Lee\cite{rf:bl91} and
Blandford et al.\cite{rf:bland91}. Gravitational influences from
large-scale structures are involved in the terms giving the Ricci and
Weyl focusing, through the density power spectrum or
the numerically simulated matter distribution. The angular correlation
functions of optical scalars were derived in connection with the
two-point density correlation function, and higher correlations also were
studied by Villumsen,\cite{rf:vill96} Bernardeau\cite{rf:bern97} and
Nakamura.\cite{rf:nakam97}
The perturbative formulation for the image deformations has been
derived independently from the null-geodesic equation
(Linder,\cite{rf:lind90} Martinez et al.,\cite{rf:mart92} Martinez and
Sanz\cite{rf:mart97}). This approach
also has been applied to studies concerning weak image deformation of
high-redshift objects and the smoothing of the CMB anisotropy
(Cayon et al.,\cite{rf:cay93} Seljak\cite{rf:selj96} and Jain and
Seljak\cite{rf:jain97}). In \S 4, a recent result obtained using the power
spectrum approach by Nakamura is given.
\section{Lensing in the multi-lens-plane method}
\subsection{Methodology}
The inhomogeneities that perturb a light bundle as it travels from
a source to an observer are of sizes much smaller than the
distance that the light traverses. This has motivated a number of
authors to use the multi-plane gravitational lensing
approach to study how light propagates in inhomogeneous universes
(e.g. Schneider and Weiss,\cite{rf:sw88a}\tocite{rf:sw88b}
Jaroszy\'nski et al.,\cite{rf:jppg}
Jaroszy\'nski,\cite{rf:jaro91a}\tocite{rf:jaro92b} Wambsganss, Cen
and Ostriker\cite{rf:wam98}).
In this approach one first idealizes the inhomogeneities as being
distributed on a series of thin lens planes which are arranged
perpendicular to the line of sight. Then one assumes that lensing
only occurs in each of those planes. This way one can analyze the
lensing properties of each plane separately and let the light beam
carry the lensing effect of each plane while propagating from one
plane to the next. For this paper to be self-contained
we give a brief review of multi-plane lensing theory, following closely
the description and notation of Schneider et al.\cite{SCH92}
\subsection{Multi-plane gravitational lensing}
\def\mib#1{\mbox{\boldmath $#1$}}
Consider $N$ lens planes located at redshifts $z_{i}$,
with $i=1,N$, and ordered such that $z_{i}<z_{j}$ for $i<j$.
Figure~1 displays an example with $N=2$. All angles are
exaggerated for clarity. Each lens plane is characterized by its
respective surface mass density $\sigma_{i}(\mib{\xi}_{i})$,
where $\mib{\xi}_{i}$ is the impact vector of the ray in the
$i$-th lens plane. Let $\hat{\mib{\alpha}}_i(\mib{\xi}_{i})$
denote the
deflection angle the light ray experiences in the $i$-th plane at a
position $\mib{\xi}_{i}$. {From} this geometry, we can derive the
{\it lens equation},
\begin{equation}\label{p1}
\mib{\eta} = \frac{D_S}{D_1}\mib{\xi}_1 - \sum_{i=1}^{N} D_{iS}
\hat{\mib{\alpha}}_i(\mib{\xi}_{i}) \,,
\end{equation}
\noindent where $\mib{\eta}$ is the
source position vector (in the source plane),
$\mib{\xi}_{i}$ is the impact vector in the $i$-th plane,
$D_j$ is the angular diameter distance between
the $j$-th plane and the observer, and $D_{ij}$ is the angular diameter
distance between the $i$-th and $j$-th planes, with $S\equiv N+1$
identifying the source plane.
Knowing the impact vector $\mib{\xi}_1$ in the image plane,
the impact vector in subsequent planes can be
obtained recursively using
\begin{equation}\label{p2}
\mib{\xi}_{j} = \frac{D_j}{D_1}\mib{\xi}_1 - \sum_{i=1}^{j-1} D_{ij}
\hat{\mib{\alpha}}_i(\mib{\xi}_{i}) \,.
\end{equation}
\noindent
The deflection angle is related to the surface density by
\begin{equation}\label{p3}
\hat{\mib{\alpha}}_i(\mib{\xi})=\frac{4G}{c^2}\int\!\!\!\int
\sigma_i(\mib{\xi}')
\frac{\mib{\xi}_i-\mib{\xi}'}{|\mib{\xi}_i-\mib{\xi}'|^2}d^2\xi'\,,
\end{equation}
\noindent
where $G$ is the gravitational constant, $c$ is the speed of light,
and the integral extends over the lens plane. We can rewrite this expression
conveniently as
\begin{eqnarray}
\label{p4}
\hat{\mib{\alpha}}_i(\mib{\xi}_i)&=&\nabla\hat\psi_i(\mib{\xi}_i)\,,\\
\label{p5}
\hat\psi_i(\mib{\xi}_i)&=&\frac{4G}{c^2}\int\!\!\!\int
\sigma_{i}(\mib{\xi}')\ln |\mib{\xi}_i-\mib{\xi}'|d^2\xi'\,.
\end{eqnarray}
\noindent It is useful to rewrite these equations
in a dimensionless form. Define
for each lens plane a critical surface density as
\begin{equation}\label{p6}
\sigma_{i,\rm cr}=\frac{c^2D_S}{4\pi GD_iD_{iS}}\,,
\end{equation}
\begin{figure}
\epsfxsize=12.cm
\centerline{\epsfbox{pfig1.ps}}
\caption{Multi-lens-planes.}
\label{fig:1}
\end{figure}
\noindent
and introduce the following dimensionless quantities :
\begin{eqnarray}
\label{p7}
{\bf \mib x}_{i}&=&\frac{\mib{\xi}_i}{D_i}\,,\qquad 1 \leq i \leq N+1 \,;\\
\label{p8} \kappa_{i}({\bf \mib x}_i)
&=&\frac{\sigma_i}{\sigma_{i,\rm cr}}\,,\qquad 1 \leq i \leq N \,.
\end{eqnarray}
\noindent Equations (\ref{p2}), (\ref{p4}) and (\ref{p5}) reduce to
\begin{eqnarray}
\label{p9}
{\bf\mib x}_{j} &=& {\bf\mib x}_{1} - \sum_{i=1}^{j-1} \beta_{ij}
\mib{\alpha}_{i}({\bf\mib x}_{i})\,,\\
\label{p10}
\mib{\alpha}_{i}({\bf\mib x}_i) &=& \nabla \psi_{i}({\bf\mib x}_i)\,,\\
\label{p11}
\psi_{i}({\bf\mib x}_{i})&=& \frac{1}{\pi}\int\!\!\!\int
\kappa_{i}({\bf\mib x}') \; \ln |{\bf\mib x}_i-{\bf\mib x}'|d^2x'\,,
\end{eqnarray}
\noindent where
\begin{equation}\label{p12}
\beta_{ij}=\frac{D_{ij}D_S}{D_jD_{iS}}\,,
\end{equation}
\noindent
and the gradient is now taken relative to ${\bf\mib x}_i$.
We can invert equation~(\ref{p11}) and obtain
\begin{equation}\label{p13}
\nabla^{2}\psi_i=2\kappa_i\,.
\end{equation}
\noindent
To compute the scaled position ${\bf\mib y}\equiv{\bf\mib x}_S$ of the source
in the source plane, we simply set $j=N+1$. Equation~(\ref{p12}) gives
$\beta_{iS}=1$, and Eq.~(\ref{p9}) becomes
\begin{equation}\label{p14}
{\bf\mib y} \equiv {\bf\mib x}_{N+1} = {\bf\mib x}_{1} - \sum_{i=1}^{N}
\mib{\alpha}_{i}({\bf\mib x}_{i})\,.
\end{equation}
\noindent
This ray-tracing equation is a mapping from the image plane
($i=1$) onto the source plane ($i=N+1$).
The effect of each lens plane on the evolution of the beam
is described by the Jacobian matrix
\begin{equation}\label{p15}
{\bf \mib A}_i({\bf \mib x}_i)
=\left(\matrix{1-\psi_{i,11} & -\psi_{i,12} \cr
-\psi_{i,21} & 1-\psi_{i,22} \cr}\right)\,,
\end{equation}
\noindent
where the commas denote differentiation
with respect to the components of ${\bf \mib x}_i$.
Since $\psi_{i,12}=\psi_{i,21}$, and Eq.~(\ref{p13}) gives
$\psi_{i,11}+\psi_{i,22}=2\kappa_i$, we can rewrite Eq.~(\ref{p15}) as
\begin{equation}\label{p16}
{\bf \mib A}_i=\left(\matrix{1-\kappa_i-S_{11} & -S_{12}\cr
-S_{12} & 1-\kappa_i+S_{11} \cr }\right)\,,
\end{equation}
\noindent where
\begin{eqnarray}
\label{p17}
S_{11}&=&\frac{1}{2}(\psi_{i,11}-\psi_{i,22})\,,\\
\label{p18}
S_{12}&=&\psi_{i,12}=\psi_{i,21}\,.
\end{eqnarray}
\noindent We now define
\begin{equation}\label{p19}
S_i=(S_{11}^2+S_{12}^2)^{1/2}\,.
\end{equation}
\noindent The determinant and trace of
${\bf \mib A}_i$ can be
expressed entirely in terms of $\kappa_i$ and $S_i$, as follows:
\begin{eqnarray}
\label{p20}
\det\,{\bf \mib A}_i&=&(1-\kappa_i)^{2}-S_i^{2}\,,\\
\label{p21}
{\rm tr}\,{\bf \mib A}_i&=&2(1-\kappa_i)\,.
\end{eqnarray}
\noindent The quantities $\mu_i\equiv1/(\det{\bf \mib A}_i)$,
$1-\kappa_i$, and
$S_i$ are called the {\it magnification}, {\it convergence}
(or Ricci focusing), and {\it shear}, respectively.
To compute the cumulative effect of all the lens planes, we consider
the Jacobian matrix of the mapping given by Eq.~(\ref{p14}),
\begin{equation}\label{p22}
{\bf\mib B}({\bf\mib x})=\frac{\partial{\bf\mib y}}{\partial{\bf\mib x}_1}
={\bf\mib I}-\sum_{i=1}^N\frac{\partial\mib{\alpha}_i}{\partial{\bf\mib x}_1}
={\bf\mib I}-\sum_{i=1}^N{\bf\mib U}_i{\bf\mib B}_i\,,
\end{equation}
\noindent where $\bf\mib I$ is the $2\times2$ identity matrix,
and ${\bf\mib U}_i$ and ${\bf\mib B}_i$ are defined by
\begin{eqnarray}
\label{p23}
{\bf\mib U}_i&=&\frac{\partial\mib{\alpha}_i}{\partial{\bf\mib x}_i}\,,\\
\label{p24}
{\bf\mib B}_i&=&\frac{\partial{\bf\mib x}_i}{\partial{\bf\mib x}_1}\,.
\end{eqnarray}
\noindent After substituting Eq.~(\ref{p10}) into Eq.~(\ref{p23}), we
obtain
\begin{equation}\label{p25}
{\bf\mib U}_i={\bf\mib I}-{\bf\mib A}_i=
\biggl(\matrix{ \psi_{i,11} & \psi_{i,12} \cr
\psi_{i,21} & \psi_{i,22} \cr }\biggr)
\,,
\end{equation}
\noindent where ${\bf\mib A}_i$ is given by Eq.~(\ref{p15}).
Hence, ${\bf\mib U}_i$ describes
the effect the $i$-th plane would have on the beam if all
the other planes were absent, and Eq.~(\ref{p22}) simply combines the
effect of all the planes. To compute the matrices ${\bf\mib B}_i$,
we differentiate Eq.~(\ref{p9}) and get
\begin{equation}\label{p26}
{\bf\mib B}_j={\bf\mib I}-\sum_{i=1}^{j-1}\beta_{ij}{\bf\mib U}_i
{\bf\mib B}_i\,.
\end{equation}
\noindent Since ${\bf\mib B}_1={\bf\mib I}$, we can use Eq.~(\ref{p26})
to compute all matricies ${\bf\mib B}_i$ by recurrence.
\subsection{The numerical algorithm}
We use the P$^3$M algorithm (Hockney and Eastwood \cite{rf:HE81}) for
the $N$-body
simulations of the large scale structure (LSS) of the universe. The
calculations evolve a system of gravitationally interacting particles in
a cubic volume with triply periodic boundary conditions, comoving with the
Hubble flow. The forces on the particles are computed by solving the
Poisson equation on a cubic grid using a Fast Fourier Transform method.
\subsubsection{Building the lens planes}
To implement the multi-plane lens method, we divide the space between
the source and the observer into a chain of cubic boxes of equal
comoving size, L$_{\rm box}$. We first need to determine the
redshifts of the interfaces between these cubic boxes. Let us assume that
the photons that reach the observer at present entered a particular
box at time $t'$ and redshift $z'$ and exited that box at time $t$ and
redshift $z$. The redshifts $z'$ and $z$ are related by (Premadi et
al.\cite{rf:prem98})
\begin{equation}\label{p27}
L_{\rm box}=\int_{t'}^{t}\big[1+z(t)\big]c\,dt\,.
\end{equation}
Using this equation, with the appropriate relation for $z(t)$,
we can find the redshifts of the interfaces. The front
side of the box closest to the the observer is, by definition,
at $z=0$. Plugging
this value into Eq.~(\ref{p27}) gives us the redshift $z'$ of
the back side of the box, which is also the redshift $z$ of the front side
of the next box. Then, by using Eq.~(\ref{p27}) recursively, we can
compute the redshifts of all the interfaces.
As the structures inside the box might evolve while the light beam travels
across it, we decide the plane onto which we project the mass
distribution
to be at redshift where the density contrast is equal to the
time-averaged density contrast of the box.
\subsubsection{The galaxy distribution}
We consider the LSS at present ($z=0$) resulting from
the P$^3$M simulations, and we design an empirical Monte-Carlo method
for locating galaxies in the computational volume, based on the
constraints that (1) galaxies should be predominantly located in the
densest regions, and (2) the resulting distribution of galaxies
should resemble the observed distribution on the sky.
Our method is the following. We divide the present
computational volume into $128^3$ cubic cells of size $1\,{\rm Mpc}^3$, and
compute the matter density $\rho$ at the center of each
cell using the same mass assignment scheme
as in the P$^3$M code. We then choose a particular density
threshold $\rho_{\rm t}$. We locate $N$ galaxies in each cell, where
$N$ is given by
\begin{equation}
N={\rm int}\biggl(\frac{\rho}{\rho_{\rm t}}\biggr)\,.
\end{equation}
\noindent The actual location of each galaxy is chosen to be
the center of the cell, plus a random offset that is of the order
of the cell size.
This eliminates any spurious effect introduced by the use of a grid.
We then experiment with various values of the density
threshold $\rho_{\rm t}$ until the total number of galaxies comes out to be
of order 40000. This gives a number density of
$\sim0.02\,{\rm galaxies}/{\rm Mpc}^3$. This method bears some
similarities with that used by
Jaroszy\'nski.\cite{rf:jaro91a}\tocite{rf:jaro92b}
Tests showed that the observed galaxy 2-point correlation function
is fairly well reproduced (Martel et al.\cite{rf:MPM98}).
The morphological type of each galaxy is determined by using a
combination of the known relations between the present distribution
of morphological
types and the surface density of galaxies along with a Monte-Carlo method.
The locations of each galaxy at higher redshifts are determined by
combining the distribution of the galaxies and that of the particles from
the P$^3$M simulation, i.e., by following the position of the nearest particle.
We adopt the galaxy models described in Jaroszy\'nski.
\cite{rf:jaro91a}\tocite{rf:jaro92b}
The projected surface density of each galaxy is given by
\begin{equation}
\sigma(r) = \cases{\displaystyle
\frac{v^2}{4G(r^2+r_c^2)^{1/2}}\,, & $r<r_{\max}$\,; \cr
0\,, & $r>r_{\max}$\, , \cr}
\end{equation}
\noindent where $r$ is the projected distance from the center.
The parameters $r_c$, $r_{\max}$ and $v$ are the core radius, maximum radius,
and the rotational velocity, respectively. These parameters are
functions of the
luminosity and morphological types of each galaxy. We assume that the
present galaxy luminosities are given by the Schechter luminosity function,
\begin{equation}
n(L)dL=\frac{n_*}{ L_*}\left(\frac{L}{L_*}\right)^{\alpha}e^{-L/L_*}dL\,,
\end{equation}
\noindent
where $n(L)$ is the number density of galaxies per unit luminosity,
$\alpha=-1.10$, $n_*=0.0156h^{3}{\rm Mpc}^{-3}$, and $L_{B*}=1.3\times
10^{10}h^{-2}L_{\odot}$, where $L_B$ is the luminosity
in the $B$ band (Efstathiou et al. \cite{rf:EEP88}). For each galaxy
we generate a luminosity $L$ according to this distribution, and we combine
it with the galaxy morphological type to determine the values of the
parameters $v$, $r_c$ and $r_{\rm max}$ for that galaxy.
\subsection{The experiments}
\subsubsection{The cosmological models}
We are currently appyling the algorithm described above to perform a large
cosmological parameter survey. We focus on the CDM model whose
density fluctuation power spectrum is described in detail in Bunn and White.
\cite{rf:BW97}
This power spectrum is characterized by 6 independent parameters :
(1) the density parameter $\Omega_{0}$; (2) the contribution $\Omega_B$ of
the baryonic matter to the density parameter; (3) the cosmological constant
$\lambda_0$; (4) the Hubble constant $H_0$; (5) the temperature
$T_{\rm CMB}$
of the cosmic microwave background radiation; and (6) the tilt $n$ of
the power spectrum. We initially set $T_{\rm CMB}=2.7 K$ and
$\Omega_B = 0.015h^{-2}$, thus reducing the dimensionality of the
parameter space to 4. The normalization of the power spectrum is often
described in terms of the rms density fluctuation $\sigma_8$ at a
scale of $8h^{-1}$Mpc, which is a function of the 6 aforementioned
parameters. We invert this relation, treating $\sigma_8$ as an
independent parameter, and the tilt $n$ as the dependent parameter.
Presently we have 43 models with various (but restricted) values of
the above first 4 parameters.
\subsubsection{The ray shooting}
In each model we propagate beams of $341^2$ light rays backward to
redshift $z_s \approx 3$. The beam has a size of 21.9 arcseconds,
and the spacing between rays is 0.064 arcseconds. To analyze the
results of the experiments, we lay down grid on the source plane,
divided into 31 x 31 cells. Cells which have more than the average
number of rays ($341^2/31^2$=121) indicate sources which are magnified.
The fraction of such cells gives us the magnification probability.
By taking the rays located in such cells, and tracing them back to the
image plane, we can study the shape of the images, and in particular,
compute the fraction of the magnified sources that have double images.
Figure~2 displays a few examples of images of a circular source.
\begin{figure}
\epsfxsize=12.cm
\centerline{\epsfbox{pfig2.ps}}
\caption{Various images of circular sources.}
\label{fig:2}
\end{figure}
\subsection{Preliminary results}
We have so far conducted between 30 and 60 ray shooting rays on all of the
models, but results of only 35 models have been analyzed.
In a previous work (Premadi et al. 1998) we tested the methodology
we have described here by studying the statistics of magnification and
shear with respect to the lens redshift for a given source redshift.
Among other things, we have found that the lensing effect is most prominent at
intermediate redshifts although the structures are more evolved at
lower redshifts. The redshift where most lensing occurs depends upon
the cosmological models.
We also show that the magnification is dominated by convergence, with shear
contributing less than one part in \rm{$10^4$}.
In the current work we study the magnification
probability, $P_m$, the probability of double-image events given a lower limit
of magnification, $P_2$, and the distribution of the separation angle in the
double image events.
\begin{figure}
\epsfxsize=18.cm
\centerline{\epsfbox{pfig3.ps}}
\caption{Magnification probability $P_m$ of magnification 1.2 and larger
as a function of the rms density fluctuation
$\sigma_8$. The numbers in parentheses indicate the values of
($\Omega_0,\lambda_0$). Open circles correspond to $H_0=65$ and
filled circles to $H_0=75$.}
\label{fig:3}
\end{figure}
\begin{figure}
\epsfxsize=18.cm
\caption{Multi-image probability $P_2$ as a function of magnification
probability $P_m$. The numbers in parentheses indicate the values of
($\Omega_0,\lambda_0$). Open circles correspond to $H_0=65$ and
circles to $H_0=75$.}
\centerline{\epsfbox{pfig4.ps}}
\label{fig:4}
\end{figure}
\begin{figure}
\epsfxsize=20.cm
\caption{The distribution of the image separation in arc seconds. The counts
are not normalized to the numbers of runs which are not the same for all
models. Top row displays results for models with $\Omega_0=1.0$,
$\lambda_0=0.0$. The middle row for models with $\Omega_0=0.2$,
$\lambda_0=0.0$, and the bottom row for models with $\Omega_0=0.2$,
$\lambda_0=0.8$.}
\centerline{\epsfbox{pfig5.ps}}
\label{fig:5}
\end{figure}
In Fig. 3, despite the still insufficient statistics, for the majority
of models, we observe a tendency for $P_m$ to decrease with increasing
$\sigma_8$, and for the cases in which we do not observe this tendency, it is
not ruled out because of the large error bars. Having an anticorrelation
between $P_m$ and $\sigma_8$ is seemingly counter-intuitive, since
$\sigma_8$ measures the amplitude of the density fluctuation, which are
responsible for lensing. The explanation is that the magnification is
caused primarily by the matter located near the beam, whereas matter located
far from the beam are responsible for the shear. In cosmological models
like CDM, structure formation proceeds hierarchically. Small structures
form first, then merge to form larger structures. A larger $\sigma_8$
implies that this hierarchical merging process is more advanced. This means
that the clusters are more massive, and possibly also denser, but there
are fewer of them, thereby enlarging voids between them. The light beam
is therefore less likely to hit or pass near any cluster, resulting in
a smaller magnification probability for larger $\sigma_8$. However, if the
beam does hit a cluster, we would expect the effect to be stronger in
models with large $\sigma_8$, as the clusters are more massive.
In Fig. 4, the plot of $P_2$ versus $P_m$ indeed shows an anticorrelation
for several
models. Another interesting result is that at fixed $\Omega_0$, $\lambda_0$,
and $\sigma_8$, the magnification probability is insensitive to the value
of $H_0$. This results from the combination of two competing effects.
On one hand, the number of lens planes between the source and the observer
decreases with increasing $H_0$. On the other hand, the total mass in
each plane is proportional to the critical density, and thus increases with
$H_0$. Hence, models with larger $H_0$ have a smaller number of more
massive lens-planes.
The histograms of the separation angle (Fig. 5) shows the presence of double
peaks in most cases. This seems to be more prominent in the models
with larger $\sigma_8$ when the other cosmological parameters are held
fixed. The high density clusters might contain more elliptical galaxies
which have smaller cores than the field galaxies. A more compact core is
known to result in a larger separation angle. Intriguingly, models with smaller
$H_0$ seem to produce longer high separation tails. The presence of
$\lambda_0$ also seems to stretch the histogram to a higher separation tail,
although by a smaller amount than that caused by high $\sigma_8$.
In addition to this,
it also increases the counts at mid-separation. This might indicate that
structures in $\lambda_0$ models have large sizes which are responsible
for midsize separation angles.
Even at this preliminary stage, the results show the different
tendencies among various cosmological models. With better statistics, we
are hopeful that the differences can be sharpened, and that these
differences will eventually enable us to make meaningful comparisons
with observational results, thereby
discriminating the cosmological models.
\section{Cosmological lensing in the direct integration method}
In this section we study the behavior of light ray bundles in
inhomogeneous model
universes, solving directly the null geodesic equation in the
(cosmological) Newtonian approximation at the stage from an epoch
$t_1$ (with redshift $z_1 = 5$) to the present epoch $t_0$.
Light rays are deflected by lens objects such as galaxies and
non-galactic clouds. Here, these clouds are mass concentrations
which are unluminous but have masses comparable with a standard
galactic mass. It is not clear at present what average mass and
size and what number density such clouds have, compared with those
of galaxies. In some previous papers (by Tomita\cite{rf:tom98a}\tocite
{rf:tom98b}), we assumed that non-galactic clouds are dominant and
have a standard galactic mass and lensing strength similar to
that of galaxies. In this paper we consider another
case, in which only galaxies are dominant lenses and all non-galactic
clouds are very weak lenses, and compare the results concerning
optical deformations in
this case with those in the previous case. In the future
observations about the shear deformation at small angles will provide
the information for the structure and lensing strength of non-galactic clouds.
\subsection{Model universes and the ray shooting}
The background model universes have the line-element
\begin{equation}
\label{eq:ba1}
ds^2 = -(1+2\varphi/c^2) c^2 dt^2
+ (1-2\varphi/c^2)a^2(t)(d\mib{x})^2/[1 + K {1 \over 4}(\mib{x})^2]^2,
\end{equation}
where $K$ is the signature of spatial curvature $(\pm 1, \ 0)$.
The normalized scale factor $S \equiv a(t)/a(t_0)$ satisfies
\begin{equation}
\label{eq:ba2}
\Bigl({dS \over d \tau} \Bigr)^2 = {1 \over S} \Bigl[\Omega_0 -
(\Omega_0+\lambda_0-1)S + \lambda_0 S^3 \Bigr],
\end{equation}
where $\tau \equiv H_0 t$ and $a_0 (\equiv a(t_0))$
is specified by the relation
$(c H_0^{-1} /a_0)^2 = 1 - \Omega_0 - \lambda_0.$
The gravitational potential $\varphi$ is described by the Poisson equation
\begin{eqnarray}
\label{eq:ba3}
a^{-2} \Delta \varphi &=& [1 - {1 \over 4}(\mib{x})^2]^2 \Bigl[{\partial^2
\varphi \over \partial \mib{x}^2} + {{1 \over 2}x^i \over [1 + K {1 \over 4}
(\mib{x})^2]^3} {\partial \varphi \over \partial x^i}\Bigr] \cr
&=& 4 \pi G \rho_B [\rho(\mib{x})/\rho_B -1],
\end{eqnarray}
where $\rho_B (=\rho_{B0}/S^3)$ is the background density and
\begin{equation}
\label{eq:ba4}
\rho_{B0} = {3 {H_0}^2 \Omega_0 \over 8\pi G} = 2.77 \times 10^{11}
\Omega_0 h^2 M_\odot \ {\rm Mpc}^{-3},
\end{equation}
where $H_0 = 100 h {\rm Mpc^{-1} \ km \ s^{-1}}$.
In our treatment the inhomogeneities are locally periodic in the
sense that the physical situation at $\mib{x}$ is the same as that at
$\mib{x} + l \mib{n}$, where the components of $\mib{n}
(=(n^1,n^2, n^3))$ are
integers. In an arbitrary periodic box with coordinate volume $l^3$,
there are $N$ particles with the same mass $m$. It is assumed that the
force at an arbitrary point is the sum of forces from $N$ particles in
the box whose center is the point in question, and that the forces from
outside the box can be neglected.
In this subsection we consider two flat models (S model and L model) with
$(\Omega_0, \lambda_0) = (1.0, 0)$ and $(0.2, 0.8)$, respectively, and
an open model (O model) with $(0.2, 0)$. The
present lengths of the boxes are
\begin{equation}
\label{eq:ba5}
L_0 \equiv a(t_0) l = 32.5 h^{-1}, \ 50 h^{-1}, \ 50 h^{-1}{\rm Mpc}
\end{equation}
for S, L and O models, respectively. The particle number is $32^3$
in all models, and thus
\begin{equation}
\label{eq:ba6}
m (= \rho_{B0} {L_0}^3/N) = 2.90, \ 2.11, \ 2.11 \times 10^{11} h^{-1} M_\odot,
\end{equation}
respectively.
The distributions of particles in these models were derived using numerical
$N$-body simulations.
The particle size (giving the lens strength) is represented by the
softening radius $r_s~(= a(t) x_s)$, which is constant. For the
particle size we consider two lens models for comparison:
\noindent {\bf Lens model 1}. All particles in the low-density models
($\Omega_0 = 0.2$) have $r_s = 20 h^{-1}$kpc, $20 \%$ of the particles in
the flat model $(1.0, 0)$ have $r_s = 20 h^{-1}$kpc, and the
remaining particles have $r_s = 500 h^{-1}$kpc.
Thus practically, particles with $\Omega_c = 0.2$ (which we call {\it compact
lens objects} in previous papers\cite{rf:tom98a}\tocite{rf:tom98b})
play the role of lens objects. Their number density is
much larger than the galactic density $\Omega_g \sim 0.02$.
\noindent {\bf Lens model 2}.
$10 \%$ of the particles in the low-density model ($\Omega_0 = 0.2$)
and $2 \%$ of the particles in the flat model $(1.0, 0)$ have
$r_s = 20 h^{-1}$kpc, while the remaining particles have
$r_s = 500 h^{-1}$kpc. Thus only galaxies corresponding to
$\Omega_g = 0.02$ play significant roles as lens objects, and the remaining
particles are regarded as diffuse clouds.
The lens strength of realistic non-galactic clouds is
intermediate between these two lens models. The question of which
model is better may be answered by observational studies involving
lens phenomena.
The time evolution of the distribution of particles was derived by
performing the $N$-body simulation in the tree-code provided by
Suto.\cite{rf:su}
The initial particle distributions were derived using Bertschinger's
software {\it COSMICS}\cite{rf:bert}
under the condition that their perturbations are given
as random fields with the spectrum of cold dark matter, their power
$n$ is 1, and their normalization is specified as the dispersion
$\sigma_8 \ = \ 0.94$ with the Hubble constant $h = 0.5$ for $(1.0,
0)$ and $h = 0.7$ for other models with $\Omega_0 = 0.2$.
Now let us consider light propagation described by the null geodesic
equation with the null condition. Here we use $T \equiv
{1 \over 2}\ln
[a(t)/a(t_1)]$ as a time variable and $T_0 \equiv (T)_{t = t_0}$. Then
we have $dS = 2 \exp [2(T- T_0)] dT$, so that
\begin{equation}
\label{eq:ba7}
c dt = R\ c_R \ [\Omega_0 +(1 -\Omega_0 -\lambda_0) S + \lambda_0
S^3]^{-1/2} {e}^{3 T} dT,
\end{equation}
where
\begin{equation}
\label{eq:ba8}
R \equiv L_0/[(1+z_1) N^{1/3}]
\end{equation}
and
\begin{equation}
\label{eq:ba9}
c_R \equiv 2(c/H_0)/[R(1+z_1)^{3/2}].
\end{equation}
The line-element is
\begin{eqnarray}
\label{eq:ba10}
ds^2/R^2 = -{c_R}^2 [\Omega_0 &+&(1 -\Omega_0 -\lambda_0) S +
\lambda_0 S^3]^{-1} {e}^{6T} (1+\alpha \phi) dT^2 \cr
&+& (1-\alpha \phi) {e}^{4T} d\mib{y}^2 /F(\mib{y})^2,
\end{eqnarray}
where \ $y^0 \equiv T, \ y^i \equiv a(t_1) x^i/R, \ \varphi \equiv
(G m/ R) \phi$, \ $R_0 \equiv R a_0/a_1 = (1+z_1) R$,
\begin{equation}
\label{eq:ba11}
\alpha \equiv {2Gm \over c^2 R} = {3 \over \pi} {\Omega_0 \over (c_R)^2}
\end{equation}
and
\begin{equation}
\label{eq:ba11a}
F \equiv 1 -{1 \over 4}(R_0H_0/c)^2 (1 -\Omega_0 -\lambda_0)
(\mib{y})^2.
\end{equation}
The equations to be solved for light rays are
\begin{equation}
\label{eq:ba12}
{dy^i \over dT} = c_R {e}^T \tilde{K}^i,
\end{equation}
\begin{eqnarray}
\label{eq:ba13}
{d\tilde{K}^i \over dT} = &-&[3 \lambda_0 e^{4(T-T_0)}+(1 -\Omega_0
-\lambda_0)] e^{2(T-T_0)} \tilde{K}^i /G(T)
+\alpha {\partial \phi
\over \partial T} \tilde{K}^i \cr
&-&\gamma {c_R}^{-1} {e}^T \Bigl[{\partial \phi /
\partial y^i}/G(T) - 2 {\partial \phi \over
\partial y^j} \tilde{K}^j \tilde{K}^i\Bigl] \cr
&+& (R_0H_0/c)^2 (1 -\Omega_0 -\lambda_0) F^{-1} c_R e^T \cr
&&\times \Bigl[-y^j \tilde{K}^j \tilde{K}^i
+ {1 \over 2}(1 +2\alpha \phi)/G(T) \Bigl] ,
\end{eqnarray}
where $\tilde{K}^i \equiv {c_R}^{-1} e^{-T} {dy^i / dT}, \
\gamma \equiv \alpha (c_R)^2$ \ and
\begin{equation}
\label{eq:ba13a}
G(T) = \Omega_0 +(1 -\Omega_0
-\lambda_0) e^{2(T-T_0)}+ \lambda_0 e^{6(T-T_0)}.
\end{equation}
The null condition is
\begin{equation}
\label{eq:ba14}
\sum_i (\tilde{K}^i)^2 = 1 + 2\alpha \phi.
\end{equation}
While in flat models the solution is straightforward, some care must
be taken in curved models.
The potential $\phi$ is given as a solution of the Poisson equation.
Because the ratio of the second term to the first term on the
right-hand side of Eq. (\ref{eq:ba3}) is $(R_0H_0/c)^2 (\mib{y})^2
[\Delta y/\vert \mib{y}\vert] \ll 1$ for $z \leq z_1$, the Poisson
equation in a box can be approximately expressed as
\begin{equation}
\label{eq:ba15}
F(\mib{y}_c)^2 \ {\partial^2 \phi \over \partial \mib{y}^2} =
4 \pi G \rho_B [\rho(\mib{x})/\rho_B -1],
\end{equation}
where we have used $F(\mib{y}) \simeq 1$ for $y \sim$ [the box size].
$F(\mib{y})$ can be approximately replaced by the central value
$F(\mib{y}_c)$ for $y \gg $ [the box size], where the index $c$ denotes the central value in the box.
For point sources with $ \rho = m \sum_n \delta(a R (\mib{y}
-{\mib{y}_n}))$ \ ($n$ is the particle number), we have
\ $\phi = \phi_1 + \phi_2$, \ where
\begin{equation}
\label{eq:ba16}
\phi_1 = - F(\mib{y}_c) e^{-2T} \sum_n {1 \over \vert \mib{y}
-\mib{y}_n\vert},
\end{equation}
and $\phi_2$ represents the contribution from the homogeneous background
density. Here let us use for $\mib{y}$ another vectors $\bar{\mib{y}}$
expressing the space in a locally flat manner, where the two coordinates
are related as
\begin{equation}
\label{eq:ba17}
\bar{\mib{y}} = \int^{\mib{y}}_{\bf 0} d\mib{y}/F(\mib{y}),
\end{equation}
and the lengths between two points in boxes in two coordinates are
approximately related as
\begin{equation}
\label{eq:ba18}
\Delta \bar{\mib{y}} = \Delta \mib{y}/F(\mib{y}_c).
\end{equation}
Then $\phi_1$ can be expressed in terms of $\bar{\mib{y}}$ in the usual
manner as
\begin{equation}
\label{eq:ba19}
\phi_1 = - e^{-2T} \sum_n {1 \over \vert \bar{\mib{y}}
-\bar{\mib{y}}_n\vert}.
\end{equation}
Corresponding forces are expressed as $f_i \equiv \partial
\phi_1/ \partial y^i = [\partial \phi_1/ \partial {\bar y}^i]/F(\mib{y}_c)$.
It should be noted that the contribution of $\partial \phi/ \partial T$ is
negligibly small, compared with that of $f_i$.
In flat models the universe is everywhere covered with periodic boxes
continuously connected as in Fig. 6. In open models it cannot be covered
in such a way, but we can consider only a set of local periodic boxes
connected along each light ray, as in Fig. 7. In these boxes we can
describe the evolution in the distribution of particles in terms of
local flat coordinates $\bar{\mib{y}}$, because the size of boxes is
much smaller than the curvature radius.
For the integration of the above null-geodesic equations, we calculate
the potential at a finite number of points on the ray which are
given at each time step ($\Delta T$). Then particles near one of the
points on the ray have a stronger
influence upon the potential than particles far from any points.
To avoid this unbalance in the calculation of the potential, we take
an average of the potential $\phi_1$ by integrating it analytically
over the interval between one of the points and the next point.
The expression for an averaged potential $\bar \phi$ was given in a
previous paper. Moreover, to take into account the finite particle
size as galaxies or non-galactic clouds,
we modify the above potential for point sources using the
softening radii $y_s = a(t_1) x_s/R$. The modified potential is
produced by replacing $(y-y_n)^2$ by $(y-y_s)^2 + (y_s)^2$ in the
potential for point sources.
The initial values of $\tilde{K}^i$ are given so as to satisfy
Eq. (\ref{eq:ba14}). The integration of Eqs. (\ref{eq:ba12}) and
(\ref{eq:ba13}) with the modified potential was performed using the Adams
method. As the time step we used $\Delta
T = [\ln 6/2]/N_s$ with $N_s = 3000$ (in most cases) $ - 10000$.
Here we consider the comoving volume of the region through which a ray
bundle with solid angle $\theta^2$ pass at the interval $0 \leq z \leq
z_i$ in the flat model $(1.0, 0)$. This volume is equal to the volume
of the cone $V_R = {1 \over 3}
\theta^2 \ [2 c H_0 (1 - 1/\sqrt{1+z_i})]^3$. Its ratio
to the volume of the periodic box $V_B \ [= (32 h^{-1}{\rm Mpc})^3]$ is
$1.2 \times 10^{-3}$ for $z_i = 5$ and $\theta = 10$arcsec. This small
ratio implies that the influence of the periodicity of the box on the
the statistics of ray bundles is negligibly small, as long as the ray
bundles are not directed in some special directions with respect to the box.
\begin{figure}[htb]
\parbox{.471\textwidth}
\epsfxsize=5.5cm
\epsfbox{tfig1.ps}
\caption{Light rays and periodic boxes in a flat space.}}
\hspace{1mm}
\parbox{.471\textwidth}{
\epsfxsize=4.3cm
\epsfbox{tfig2.ps}
\caption{Light rays and periodic boxes in an open space.}}
\end{figure}
\subsection{Statistical behavior of optical scalars}
We treat the deformation of ray bundles over the interval from $z = 0$
to $z = 5$ measured by an observer in a periodic box. Here we consider
the ray bundles reaching the observer (or emitted backwards) in a
regular form such that
the rays are put in the same separation angle $\theta$, and
calculate the relative change in the angular positions of the rays
which increases
with redshift in the past direction. From this change we find the
behavior of optical scalars.
Basic ray bundles consist of 5 $\times$ 5 rays that are put in a square
form with the same
separation angle $\theta = 2 - 360$ arcsec. Many bundles coming
from all directions in
the sky are considered. Here we take 200 bundles coming from randomly
chosen directions for each separation angle. In order to express
relative angular positions of rays, we use two orthogonal vectors,
$e^i_{(1)}$ and $e^i_{(2)}$, in the plane perpendicular to the first
(fiducial) background ray vector $(\tilde{K}^i)_B$.
Then the angular coordinates $[X(m,n), Y(m,n)]$ of 25 rays relative to
the first (fiducial) ray with $(m,n) = (1,1)$ at any epoch are defined by
\begin{eqnarray}
\label{eq:st2}
X(m,n) &=& \sum_i [y^i(m,n) - y^i(1,1)]\ e^i_{(1)}/y_B (1,1) +X(1,1),\cr
Y(m,n) &=& \sum_i [y^i(m,n) - y^i(1,1)]\ e^i_{(2)}/y_B (1,1) +Y(1,1),
\end{eqnarray}
where $y^i = y^i_B + \delta y^i$ and $y_B = [\sum_i (y^i_B)^2]$.
Since all angular intervals of rays at the observer's point are the same ($=
\theta$), the differentiation of angular coordinates of the
rays at any epoch with respect to those at the observer's points is given
by the differences
\begin{eqnarray}
\label{eq:st3}
A_{11}(m,n) &=& [X(m+1,n) - X(m,n)]/\theta, \cr
A_{12}(m,n) &=& [X(m,n+1) - X(m,n)]/\theta, \cr
A_{21}(m,n) &=& [Y(m+1,n) - Y(m,n)]/\theta, \cr
A_{22}(m,n) &=& [Y(m,n+1) - Y(m,n)]/\theta,
\end{eqnarray}
where $m$ and $n$ run from 1 to 5. From the matrix $A_{ij} (m,n)$ we
derive the optical scalars in the standard manner,\cite{SCH92}
as the convergence ($\kappa (m,n)$),
the shear ($\gamma_i (m,n), \ i=1,2$), and the amplification ($\mu
(m,n)$) defined by
\begin{eqnarray}
\label{eq:st4}
\kappa (m,n) &=& 1 - {\rm tr}(m,n)/2, \quad \gamma_1 (m,n) = [A_{22}(m,n) -
A_{11}(m,n)]/2, \cr
\gamma_2 (m,n) &=& -[A_{12}(m,n) + A_{21}(m,n)]/2, \cr
\gamma^2 &\equiv& (\gamma_1)^2 + (\gamma_2)^2 = [{\rm tr}(m,n)]^2 -
\det(A_{ij}(m,n)), \cr
\mu (m,n) &=& 1/\det(A_{ij}(m,n)),
\end{eqnarray}
where the trace is ${\rm tr}(m,n) = A_{11}(m,n) + A_{22}(m,n)$.
The average optical quantities in each bundle are defined as the
averages of optical quantities for all rays in the bundle as
\begin{equation}
\label{eq:st5}
\bar{\kappa} = \Bigl[\sum_m \sum_n \kappa (m,n)\Bigr]/4^2, \quad
\bar{\kappa^2} = \Bigl[\sum_m \sum_n (\kappa (m,n))^2\Big]/4^2,
\end{equation}
and so on. In this averaging process the contributions from smaller
scales can be cancelled and smoothed-out. The above optical scalars
at the separation angle $\theta$ are accordingly derived in the
coarse-graining on this smoothing scale.
For the present statistical analysis we excluded caustic cases,
considering only the case of {\it weak lensing} in the sense of
{\it no caustics}.
The averaging for all non-caustic ray bundles is denoted using
$\langle \rangle$
as $\langle\kappa^2\rangle, \ \langle\gamma^2\rangle$ and
$\langle(\mu -1)^2\rangle$. Because
$\kappa(m,n), \ \gamma_i (m,n)$ and $\mu (m,n) -1$ take positive and
negative values with almost equal frequency, $\langle\kappa\rangle,
\ \langle\gamma_i\rangle$
and $\langle\mu\rangle -1$ are small.
In Figs. 8 and 9, we show the behavior of $\langle\gamma^2\rangle$ for various
separation angles $\theta = 2$ arcsec $- \ 360$ arcsec (= $6$ arcmin)
at epochs $z = 1.0$ and $2.0$, respectively. Similarly in Figs. 10 and
11, we show the behavior of $\langle(1 - \mu)^2\rangle$ at epochs
$z = 1.0$ and
$2.0$, respectively. The calculations were
performed at $\theta = 2, 5, 10, 30, 60,$ and $360$ arcsec.
The behavior of $\langle\kappa^2\rangle$ is similar to that of $\langle
\gamma^2\rangle$,
as was shown in previous papers. In all figures, the values in the two
lens models are shown using solid and dotted lines.
It is found in the low-density models that the ratios of $\langle
\kappa^2\rangle^{1/2}$,
$\langle\gamma^2\rangle^{1/2}$ and $\langle(1 - \mu)^2
\rangle^{1/2}$ in the lens model 1
to those in the lens model 2 are about 4 at the separation angle $\theta
= 2$ arcsec. The ratios decrease with the increase of $\theta$, and are
$< 2$ at $\theta = 360$ arcsec. That is, the remarkable difference between
the two lens
models appears at the small angles $\theta \simeq 2$ arcsec.
The values of optical scalars for $\theta \sim 360$ arcsec are roughly
consistent with the results of Bernardeau et al. (cf. their Figs.
3 and 4)\cite{rf:bern97} and Nakamura\cite{rf:nakam97} in their treatments.
At all separation angles and for both lens models, $\langle\kappa^2\rangle,
\langle\gamma^2\rangle$ and $\langle(1 - \mu)^2\rangle$ in the open
model ($\Omega_0 = 0.2$)
are larger than those in the flat model (L). In model S they
are largest at all angles for the lens model 2 and at large angles
($\theta \sim 360$ arcsec) in the lens model 1, but they are smallest
at small angles ($\theta < 30$ arcsec) because the role of
non-galactic clouds on small-scale is small.
In Tables I, II and III, we show for models S, L and O the
numerical values of $\langle\kappa\rangle,\
\langle\kappa^2\rangle^{1/2}, \ \langle\mu\rangle, \ \langle(\mu -1)^2
\rangle^{1/2}, \ \langle\gamma_1\rangle$ and
$\langle\gamma^2\rangle^{1/2}$ \ at epochs $z = 1,2, .., 5$ for $\theta = 2$
arcsec.
The influence of lensing on source magnitudes is
\begin{equation}
\label{eq:st6}
\Delta m = {5 \over 2} \log [1 + \langle(1-\mu)^2\rangle^{1/2}] \simeq 1.09
\langle(1-\mu)^2\rangle^{1/2}.
\end{equation}
If $\theta = 2$ arcsec, the value of $\Delta m$ is $(0.058, 0.034),
(0.077, 0.019)$ and $(0.133, 0.022)$ for the
lens model $(1, 2)$, respectively, at $z = 1$ in models S, L and O.
If the lens model 1 is more realistic, therefore, this lens correction may
introduce some sensitive modification to the standard selection of cosmological
models in the $[m, z]$ relation. Here it should be noted that the
lensing effect at small angles, which is caused by small-scale
nonlinear inhomogeneities, is important for the magnitude correction
of small high-redshift objects, such as SNe Ia, in cosmological
observations.\cite{rf:perl}\tocite{rf:riess}
\begin{figure}
\epsfxsize=7.5cm
\centerline{\epsfbox{tfig3.ps}}
\caption{The angular dependence of $\langle\gamma^2\rangle^{1/2}$.
Solid and dotted lines represent behavior for lens models 1 and 2,
respectively, at $z = 1$.
The separation angle $\theta$ is in units of arcsec.}
\label{fig:8}
\end{figure}
\begin{figure}
\epsfxsize=7.5cm
\centerline{\epsfbox{tfig4.ps}}
\caption{The angular dependence of $\langle\gamma^2\rangle^{1/2}$.
Solid and dotted lines denote behavior for lens models 1 and 2,
respectively, at $z = 2$.
The separation angle $\theta$ is in the unit of arcsec.}
\label{fig:9}
\end{figure}
\begin{figure}
\epsfxsize=7.5cm
\centerline{\epsfbox{tfig5.ps}}
\caption{The angular dependence of $\langle(1 - \mu)^2\rangle^{1/2}$.
Solid and dotted lines represent behavior for lens models 1 and 2,
respectively, at $z = 1$.
The separation angle $\theta$ is in units of arcsec.}
\label{fig:10}
\end{figure}
\begin{figure}
\epsfxsize=7.5cm
\centerline{\epsfbox{tfig6.ps}}
\caption{The angular dependence of $\langle(1 - \mu)^2\rangle^{1/2}$.
Solid and dotted lines represent behavior for lens models 1 and 2,
respectively, at $z = 2$.
The separation angle $\theta$ is in the unit of arcsec.}
\label{fig:11}
\end{figure}
\begin{table}
\caption{Optical quantities in model S at $\theta = 2$ arcsec.}
\label{table:1}
\begin{center}
\begin{tabular}{ccccccrc} \hline \hline
$lens$&$z$&$\langle\kappa\rangle$ & $\langle\kappa^2\rangle^{1/2}$ &
$\langle\mu\rangle$ & $\langle(1-\mu)^2\rangle^{1/2}$
& $\langle\gamma_1\rangle$ & $\langle\gamma^2\rangle^{1/2}$
\\ \hline
&1& $ 0.0005$ & $0.0245$ & 1.0036 & 0.0529 & $-0.0032$ & 0.0251\\
&2& $ 0.0004$ & $0.0398$ & 1.0075 & 0.0858 & $-0.0061$ & 0.0409\\
1&3& $ 0.0021$ & $0.0495$ & 1.0149 & 0.1084 & $-0.0080$ & 0.0514\\
&4& $ 0.0034$ & $0.0563$ & 1.0208 & 0.1251 & $-0.0096$ & 0.0593\\
&5& $ 0.0042$ & $0.0617$ & 1.0253 & 0.1390 & $-0.0109$ & 0.0653\\ \hline
&1& $ 0.0005$ & $0.0155$ & 1.0020 & 0.0315 & $-0.0026$ & 0.0152\\
&2& $ 0.0020$ & $0.0251$ & 1.0065 & 0.0516 & $-0.0049$ & 0.0238\\
2&3& $ 0.0030$ & $0.0305$ & 1.0096 & 0.0636 & $-0.0053$ & 0.0283\\
&4& $ 0.0037$ & $0.0343$ & 1.0120 & 0.0721 & $-0.0055$ & 0.0314\\
&5& $ 0.0041$ & $0.0370$ & 1.0136 & 0.0784 & $-0.0055$ & 0.0339\\ \hline
\end{tabular}
\end{center}
\end{table}
\begin{table}
\caption{Optical quantities in model L at $\theta = 2$ arcsec. }
\label{table:2}
\begin{center}
\begin{tabular}{ccccccrc} \hline \hline
$lens$&$z$&$\langle\kappa\rangle$ & $\langle\kappa^2\rangle^{1/2}$ &
$\langle\mu\rangle$ & $\langle(1-\mu)^2\rangle^{1/2}$
& $\langle\gamma_1\rangle$ & $\langle\gamma^2\rangle^{1/2}$
\\ \hline
&1& $-0.0011$ & $0.0324$ & 1.0025 & 0.0705 & 0.0008 & 0.0347\\
&2& $-0.0036$ & $0.0675$ & 1.0134 & 0.1572 & 0.0018 & 0.0714\\
1&3& $-0.0045$ & $0.0974$ & 1.0345 & 0.2407 & 0.0020 & 0.1002\\
&4& $-0.0065$ & $0.1216$ & 1.0560 & 0.3209 & 0.0023 & 0.1214\\
&5& $-0.0070$ & $0.1405$ & 1.0805 & 0.4000 & 0.0030 & 0.1372\\ \hline
&1& $ 0.0001$ & $0.0084$ & 1.0004 & 0.0174 & $-0.0004$ & 0.0088\\
&2& $ 0.0004$ & $0.0175$ & 1.0021 & 0.0367 & $-0.0018$ & 0.0180\\
2&3& $ 0.0001$ & $0.0239$ & 1.0025 & 0.0505 & $-0.0020$ & 0.0245\\
&4& $ 0.0001$ & $0.0291$ & 1.0038 & 0.0618 & $-0.0019$ & 0.0296\\
&5& $ 0.0005$ & $0.0333$ & 1.0055 & 0.0710 & $-0.0017$ & 0.0334\\ \hline
\end{tabular}
\end{center}
\end{table}
\begin{table}
\caption{Optical quantities in model O at $\theta = 2$ arcsec. }
\label{table:3}
\begin{center}
\begin{tabular}{ccccccrc} \hline \hline
$lens$&$z$&$\langle\kappa\rangle$ & $\langle\kappa^2\rangle^{1/2}$
& $\langle\mu\rangle$ & $\langle(1-\mu)^2\rangle^{1/2}$
& $\langle\gamma_1\rangle$ & $\langle\gamma^2\rangle^{1/2}$
\\ \hline
&1& $-0.0002$ & $0.0425$ & 1.0221 & 0.1219 & $-0.0028$ & 0.0426\\
&2& $-0.0054$ & $0.0796$ & 0.9961 & 0.2210 & $-0.0041$ & 0.0832\\
1&3& $-0.0069$ & $0.1157$ & 1.0368 & 0.6234 & $ 0.0057$ & 0.1225\\
&4& $-0.0081$ & $0.1519$ & 1.0790 & 0.5638 & $ 0.0078$ & 0.1602\\
&5& $-0.0064$ & $0.1872$ & 1.1627 & 1.1880 & $ 0.0120$ & 0.1980\\ \hline
&1& $-0.0003$ & $0.0099$ & 0.9997 & 0.0205 & $-0.0005$ & 0.0120\\
&2& $-0.0012$ & $0.0209$ & 0.9993 & 0.0441 & $-0.0015$ & 0.0209\\
2&3& $ 0.0000$ & $0.0314$ & 1.0041 & 0.0682 & $-0.0035$ & 0.0284\\
&4& $ 0.0010$ & $0.0410$ & 1.0089 & 0.0915 & $-0.0047$ & 0.0370\\
&5& $ 0.0021$ & $0.0496$ & 1.0144 & 0.1140 & $-0.0059$ & 0.0451\\ \hline
\end{tabular}
\end{center}
\end{table}
For the statistical analysis we used 200 ray bundles reaching an
observer in a single inhomogeneous model universe. This number of ray
bundles may be too small to cover the influences from complicated
inhomogeneities in all directions. To obtain more robust statistical
results it may be necessary to use more ray bundles and more model
universes produced with random numbers.
Finally, we touch upon recent observations of cosmological shear due to weak
lensing and their relation to our results. Fort et al.\cite{rf:fort}
attempted
measurements of a coherent shear from foreground mass condensations in
the fields of several luminous radio sources. Schneider et
al. \cite{rf:sch}
determined the shear in the field ($2$ min $\times 2$ min)
containing a radio source PKS1508-05 with $z = 1.2$, and their result
is that the shear is about 0.03 for an angular scale of 1 min.
This value may be consistent with the low-density models in the lens
model 1, but more observational data are necessary to deduce any
conclusions.
In order to clarify the lensing strength of
non-galactic clouds it is important to have observational values
of the shear at small angles $\theta \approx 2$ arcsec.
\newcommand{\bm}[1]{\mbox{\boldmath $#1$}}
\section{Analytic evaluation of cosmic shear through the power spectrum
approach \protect\footnote{In this section we use units in which $c=H_0=1$
and follow Misner et al.\protect\cite{MIS73} as regards the sign convention
of curvature tensors.} }\label{sec:ttn}
In this section we analytically estimate the shear and the image
amplification caused by large scale density inhomogeneity in the
universe.{\cite{rf:gunnb}~\cite{rf:bl91}\tocite{rf:nakam97}
emanating from us to the past, and let $\chi$ be the affine parameter
along the central ray in the bundle and $k^a = (d/d\chi)^a$ be its
tangent vector. Deformation in the cross section of the bundle is
described by optical scalars, the expansion $\theta = \frac12
\nabla_{\!\!a} k^a$ and shear $\sigma = \frac12 \epsilon^a \epsilon^b
\nabla_{\!\!a} k_b$, where $\epsilon^a = e_1^a + i e_2^a$ is the complex
dyad basis on the cross section. The rotation $\omega = \frac12 {\rm
Im} ( \epsilon^{*a} \epsilon^b ) \nabla_{\!\!a} k_b$ identically
vanishes for a bundle which emanates from a single point, so we set
$\omega=0$ in the following. These evolve along the ray according to the
equations \cite{SCH92,SAS93}
\begin{equation}\label{n1}
\dot\theta + \theta^2 + |\sigma|^2 = -{\cal R} \,,\quad \dot\sigma +
2\theta\sigma = -{\cal F} \,,
\end{equation}
where the overdot denotes $d/d\chi$, and the sources of deformation are
${\cal R} = \frac12 R_{ab} k^a k^b$ and ${\cal F} = \frac12 C_{acbd}
\epsilon^a \epsilon^b k^c k^d$. The observable effect of this
deformation is the evolution of a deviation vector $\xi^a$ connecting
two nearby rays on the cross section:
\begin{equation}
\dot \xi = \theta\xi + \sigma \xi^*.
\end{equation}
where $\xi=\epsilon_a\xi^a$. This deviation vector is written as a
linear mapping from the initial value of $\dot\xi$ at $\chi=0$:
\begin{equation}\label{n3}
\xi = \Lambda^* \dot\xi_0 + \Gamma \dot \xi_0^*.
\end{equation}
Here $\dot\xi_0$ represents the angular separation between the
nearby rays in the observer's sky, and ${\rm Re}\Lambda$, ${\rm
Im}\Lambda$ and $\Gamma$ represent respectively the degrees of focusing,
rotation and shear in the shape of the cross section as seen from the
observer. From equation (\ref{n1}) it follows that
\begin{equation}\label{n4}
\dot\Lambda = \theta\Lambda + \sigma^* \Gamma \,,\quad
\ddot\Lambda = -({\cal R}\Lambda + {\cal F}^*\Gamma ),
\end{equation}
\begin{equation}\label{n5}
\dot\Gamma = \sigma \Lambda + \theta\Gamma \,,\quad
\ddot\Gamma = -({\cal F}\Lambda + {\cal R} \Gamma).
\end{equation}
The angular diameter distance $D$ is defined to be (cross sectional
area/solid angle in the sky)$^{1/2}$ so $D = (|\Lambda|^2 -
|\Gamma|^2)^{1/2}$ and satisfies
\begin{equation}
\dot D = \theta D \,,\quad \ddot D = -({\cal R} + |\sigma|^2) D.
\end{equation}
Initial conditions for these differential equations at $\chi=0$ are $0=
\Lambda= \Gamma= \dot\Gamma= D= \sigma$ and $\dot\Lambda= \dot D =1$.
The following form of the metric is used to describe the globally
homogeneous but locally inhomogenous universe
\begin{equation}\label{n7}
d\hat s^2 = a^2 ds^2 = a^2(\eta)\,[ -(1+2\Phi) d\eta^2 + (1-2\Phi)
\gamma_{ij} dx^i dx^j ].
\end{equation}
Here $a$ is the scale factor and $\gamma_{ij}$ is the 3-metric on the
constant time hypersurface of curvature $K=\Omega_0+\lambda_0-1$. The
gravitational potential $\Phi$ ($\ll 1$) incorporates the local
inhomogeneity generated by the density contrast $\delta =
(\rho-\bar\rho)/\bar\rho$ and satisfies the Poisson equation
\begin{equation}\label{n8}
\nabla^2 \Phi = \frac32\Omega_0 \frac\delta a,
\end{equation}
where $\nabla^2$ is the Laplacian associated with $\gamma_{ij}$ (we only
consider a sub-horizon scale inhomogeneity whose wavelength is much
smaller than $|K|^{-1/2}$). Since the null geodesics are unaffected by
the conformal transformation, we work in the scaled metric $ds^2$ in
what follows. Accordingly the ``distances'' $\hat\Lambda$ and
$\hat\Gamma$ (Eq.(\ref{n3})) in the metric $d\hat s^2$ are related to
those in $ds^2$ as $\hat\Lambda = a \Lambda$ and $\hat\Gamma =
a\Lambda$. The affine parameter $\chi$ is related to the redshift $z=1/a
-1$ as
\begin{equation}\label{n9}
\chi(z) = \int_a^1 dx\, (\Omega_0 x - K x^2 + \lambda_0 x^4)^{-1/2}.
\end{equation}
From Eq. (\ref{n7}) we obtain ${\cal R} = K + \nabla^2 \Phi$ and
${\cal F} = \epsilon^i\epsilon^j \nabla_{\!\!i}\nabla_{\!\!j} \Phi$. We
regard $\delta\!{\cal R}=\nabla^2\Phi$ and $\delta\!{\cal
F}=\epsilon^i\epsilon^j \nabla_{\!\!i}\nabla_{\!\!j} \Phi$ as
perturbations to $\bar{\cal R}=K$ and $\bar{\cal F}=0$, and solve
Eqs. (\ref{n4}) and (\ref{n5}) peturbatively. The zeroth order
solution in an exactly homogenous universe is
\begin{equation}\label{n10}
\bar\Lambda(\chi) = \bar D(\chi) = \left\{
\begin{array}{ll}
K^{-1/2} \sin(K^{1/2}\chi), & (K>0) \\
\chi, & (K=0) \\
(-K)^{-1/2} \sinh[(-K)^{1/2}\chi], & (K<0)
\end{array} \right.
\end{equation}
and $\bar\Gamma(\chi)=0$, and hence the image remains undistorted with only an
isotropic focusing. The differential equation for the first order
quantities $\delta\!\Lambda= \Lambda - \bar\Lambda$ and $\delta\!\Gamma=
\Gamma - \bar\Gamma$ are
\begin{equation}
\ddot{\delta\!\Lambda} = -(K\,\delta\!\Lambda
+ \bar D\,\delta\!{\cal R}) \,,\quad \ddot{\delta\!\Gamma} =
-(K\,\delta\!\Gamma + \bar D\,\delta\!{\cal F}),
\end{equation}
which are integrated with the initial conditions $0= \delta\!\Lambda=
\delta\!\Gamma= \dot{\delta\!\Lambda}= \dot{\delta\!\Gamma}$ as
\begin{equation}\label{n12}
\delta\!\Lambda(\chi) = -\int_0^\chi d\chi' \, \delta\!{\cal R}(\chi')
\, \bar D(\chi') \,\bar D(\chi-\chi'),
\end{equation}
\begin{equation}\label{n13}
\delta\!\Gamma(\chi) = -\int_0^\chi d\chi' \, \delta\!{\cal F}(\chi')
\, \bar D(\chi') \,\bar D(\chi-\chi').
\end{equation}
Note that $\delta\!\Lambda$ is real, so that there is no rotation
in the image to this order. Let us define
\begin{equation}
\kappa = -\delta\!\Lambda/\bar D \,,\quad \gamma = \delta\!\Gamma/\bar D.
\end{equation}
The amplification of the brightness of images relative to that in a
homogenous universe is given by $(\bar D/D)^2=1+2\kappa$ to first order,
so $\kappa$ (convergence) measures the fluctuation in the image
brightness due to the lensing by the density inhomogeneity. Similiarly
$\gamma$ (shear) is a dimensionless measure of the image distortion due
to the lensing.
The sources $\delta\!{\cal R}$ and $\delta\!{\cal F}$ of the image
distortion are stochastic variables which average to zero: $\langle
\delta\!{\cal R}\rangle = \langle \delta\!{\cal F}\rangle =0$. To
evaluate the statistical properties of this distortion, it is
convenient to Fourier-transform the density field at time $\eta$ as
$\tilde\delta(\bm k,\eta)= \int d^3\bm x\, \delta(\bm
x,\eta)\,\exp(-i\bm k\cdot \bm x)$ (valid for $k^2\gg |K|$) and define
the power spectrum $P(k,\eta)$ as $\langle \tilde\delta (\bm k,\eta)\,
\tilde\delta (\bm k',\eta) \rangle = (2\pi)^3 P(k,\eta)\,
\delta_{\rm\scriptscriptstyle D} (\bm k- \bm k')$. Then from Eq.
(\ref{n8}) the two point correlation of $\delta\!{\cal R}$ at two
locations $\bm x$ and $\bm x'$ is
\begin{equation}
\langle\delta\!{\cal R}(\bm x)\, \delta\!{\cal R}(\bm x') \rangle =
\frac94 \Omega_0^2 \int
\frac{d^3\bm k}{(2\pi)^3} \,\frac{P(k,\eta)}{a^2(\eta)} \,
\exp[i\bm k\cdot (\bm x-\bm x')].
\end{equation}
Below we calculate the two point correlations of convergence
$\langle\kappa_{\rm\scriptscriptstyle A}\, \kappa_{\rm\scriptscriptstyle
B} \rangle$ and shear $\langle\gamma_{\rm\scriptscriptstyle A}\,
\gamma_{\rm\scriptscriptstyle B}^* \rangle$ between images A and B at
the same redshifts separated by a small angle vector $\bm\theta$ on the
sky. Since the integrals in Eqs. (\ref{n12}) and (\ref{n13}) can
be performed along the unperturbed path to first order, $\bm k\cdot
(\bm x-\bm x')$ in the exponent is written as $\bm
k_{\scriptscriptstyle\perp}\cdot \bm\theta \bar D(\chi) +
k_{\scriptscriptstyle\parallel} (\chi-\chi')$ where $\bm
k_{\scriptscriptstyle\perp}$ (or $k_{\scriptscriptstyle\parallel}$) is
the wavenumber perpendicular (or parallel) to the line of sight. When
$\theta\ll 1$ we can approximate $k_{\scriptscriptstyle\parallel} \ll
k_{\scriptscriptstyle\perp}$ and obtain
\begin{equation}\label{n16}
\langle\delta\!{\cal R}(\bm x)\, \delta\!{\cal R}(\bm x') \rangle
= \frac{9\pi}4 \Omega_0^2 \, \delta_{\rm\scriptscriptstyle D}
(\chi-\chi') \int_0^\infty \frac{dk}k\,\frac{\Delta^2(k,\eta)}
{k\, a^2(\eta)} \,J_0[k\theta\bar D(\chi)],
\end{equation}
where $\Delta^2(k,\eta) = k^3\,P(k,\eta)/(2\pi^2)$. A similar
calculation shows that $\langle \delta\!{\cal F}(\bm x)\, \delta\!{\cal
F}^*(\bm x') \rangle = \langle \delta\!{\cal R}(\bm x)\, \delta\!{\cal
R}(\bm x') \rangle$ and $\langle \delta\!{\cal R}(\bm x)\, \delta\!{\cal
F}(\bm x') \rangle =0$. From Eqs. (\ref{n12}), (\ref{n13}) and
(\ref{n16}) there results \footnote{ In conventional units, Eq.
(\ref{n17}) should be divided by $(c/H_0)^4$.}
\begin{eqnarray}
C(\theta) &=& \langle \kappa_{\rm\scriptscriptstyle A}\,
\kappa_{\rm\scriptscriptstyle B} \rangle
= \langle \gamma_{\rm\scriptscriptstyle A} \,
\gamma_{\rm\scriptscriptstyle B}^* \rangle \nonumber \\
&=& \frac{9\pi}4 \Omega_0^2 \int_0^\chi d\chi'\,
\frac{\bar D^2(\chi') \bar D^2(\chi-\chi')}{\bar D^2(\chi)} \nonumber \\
&&\times \int_0^\infty \frac{dk}k \,\frac{\Delta^2(k,\chi')}{k\, a^2(\chi')}\,
J_0[k\theta\bar D(\chi')], \label{n17}
\end{eqnarray}
where $a(\chi')$ is given as Eq. (\ref{n9}).
\begin{figure}[htb]
\epsfxsize=8cm \centerline{\epsfbox{nfig1.eps}} \caption{Rms convergence
and shear $C^{1/2}(0)=\langle \kappa^2 \rangle^{1/2} = \langle
|\gamma|^2 \rangle^{1/2}$ as a function of the source redshift. The solid, dashed
and dotted curves correspond to $(\Omega_0, \lambda_0, \sigma_8)=
(1,0,0.5)$, $(0.3,0,1)$, $(0.3,0.7,1)$.} \label{nfig1}
\end{figure}
\begin{figure}[htb]
\epsfxsize=8cm \centerline{\epsfbox{nfig2.eps}} \caption{Two point
correlation $C^{1/2}(\theta)=\langle \kappa_{\rm\scriptscriptstyle A}\,
\kappa_{\rm\scriptscriptstyle B} \rangle^{1/2} = \langle
\gamma_{\rm\scriptscriptstyle A} \, \gamma_{\rm\scriptscriptstyle B}^*
\rangle^{1/2}$ of convergence and shear between images A and B at
redshift $z=2$ as a function of the separation angle $\theta$ in the sky. The
solid, dashed and dotted curves correspond to $(\Omega_0, \lambda_0,
\sigma_8)= (1,0,0.5)$, $(0.3,0,1)$, $(0.3,0.7,1)$.} \label{nfig2}
\end{figure}
In Fig.~\ref{nfig1}\ the root-mean-square convergence and shear
$C^{1/2}(0)=\langle \kappa^2 \rangle^{1/2} = \langle |\gamma|^2
\rangle^{1/2}$ is plotted versus the source redshift, and in Fig.
\ref{nfig2} the square root of the two point correlation
$C^{1/2}(\theta)=\langle \kappa_{\rm\scriptscriptstyle A}\,
\kappa_{\rm\scriptscriptstyle B} \rangle^{1/2} = \langle
\gamma_{\rm\scriptscriptstyle A} \, \gamma_{\rm\scriptscriptstyle B}^*
\rangle^{1/2}$ of convergence and shear between images A and B at
redshift $z=2$ is plotted versus the separation angle $\theta$ on the
sky. We use the non-linear power spectrum with the shape parameter
$\Omega_0 h=0.25$, \cite{PEA94} which reproduces well the nearby galaxy
survey data. The solid, dashed and dotted curves in the figures
correspond respectively to the cosmological models with $(\Omega_0,
\lambda_0, \sigma_8)=(1,0,0.5)$, $(0.3,0,1)$, $(0.3,0.7,1)$ (standard,
open and cosmological-constant models), where the normalizations
$\sigma_8$ of the power spectrum on the $8{\rm Mpc}/h$ scale are chosen
such that the observed abundance of low-redshift clusters is consistent
with the prediction of Press--Schechter theory.\cite{KIT97} The figures
represent the result for an infinitesimal source size because we do not
smooth the density field; for a finite source the density field should
be smoothed on the scale of the source size.
Since we are using a power spectrum well-constrained by the
observational data at $z\sim 0$, the differences between the curves in
the figures purely reflect the different behavior of backward light
propagation into the past in these cosmological models. Intuitively, the
rms amplification and shear should be larger in higher density
universes, as explained by the overall factor $\Omega_0^2$ in Eq.
(\ref{n17}). However Fig.~\ref{nfig1} \ shows that this effect is
overwhelmed by the evolutionary effect of the density inhomogeneity: the
growth of the density contrast in the standard (or open) model is
fastest (or slowest) among the three models according to the
gravitational instability picture.\cite{PEE80} Since a fast growth
implies that the inhomogeneity is rapidly erased as one goes to the past,
the lensing effect in the standard (or open) model is smallest (or
largest) in Fig.~\ref{nfig1}. Therefore the detection of cosmic shear
should allow us to probe the growth of the large-scale density
inhomogeneity, when combined with observational constraints for $z\sim
0$. On the other hand, Fig.~\ref{nfig2} \ shows that in the open (or
standard) model the decrease of correlation with increasing angle is
fastest (or slowest). This is due to the cosmological geometry
(Eqs.(\ref{n9}) and (\ref{n10})): for fixed $\theta$ and $z$ the
physical separation $\theta \bar D$ in the open (or standard) model is
largest (or smallest) among the three models. Thus the slope of the
correlation function of cosmic shear, $dC/d\ln\theta$, should be a
measure of the geometry of our universe.
\section{Concluding remarks}
In \S 2 it was shown that the multi-lens-plane method can treat many
ray bundles and realistic galaxy lens models in very useful and
practical ways to
derive statistical behavior of strong and weak lensing owing to the
simplicity of the calculations. The discussion in \S 3 shows that the
recent progress in
supercomputers has made it possible to use the direct integration methods for
lensing analyses, and that non-galactic lens objects on galactic scales (if
any) may have more important influences on the cosmological lensing
(than galactic lenses) on small-angle scales ($< 1$ min)
depending on the lensing
models. In \S 4 it was shown that perturbative methods based on the CDM
power spectrum are very useful for analyses of weak lensing on
large-angle scales ($\gg 1$ min).
For the lensing correction to the luminosity of small objects such as
super novae we must treat ray bundles with small separation angles
($\leq 2$ arcsec) for which the deflection due to the small-scale
matter distribution on galactic scales is important.
>From the analysis in \S 3 it is found that the correction can be
larger by a factor of $\sim 5$ than that due to the perturbative
methods.\cite{rf:fri}
\section*{Acknowledgements}
K.T. would like to thank Y.~Suto for helpful discussions about
$N$-body simulations. His numerical computations were performed on
the YITP computer system. \ P.W.P \ is indebted to H.~Martel for his
help in preparing the manuscript, \ to \ T. Futamase for his
valuable advice, and to the JSPS for a postdoctoral fellowship.
Her computational works was performed on the University
of Texas High Performance Computing Facility funded through R.~Matzner.
\bigskip
|
\section{Introduction}
The question whether neutrinos have non-zero mass has been an
outstanding issue in particle physics for many decades.
Recently there have been new exciting developments in the indirect
search for neutrino masses via neutrino oscillations. Major progress
has been achieved by studying atmospheric neutrinos,
culminating in the announcement of evidence for muon neutrino
oscillations by the Super-Kamiokande collaboration
\cite{atmospheric}. The most
striking signal presented in \cite{atmospheric}
is the up-down asymmetry of the
atmospheric muon neutrino flux. The
choice of this particular quantity eliminates many theoretical
uncertainties and the final result is very robust.
In fact, at present time, this result represents perhaps the best
evidence for physics beyond the Standard Model.
Another very active area of research is the study of neutrinos
coming from the Sun. Ever since the Homestake experiment \cite{Cl}
reported its first
results, there has been disagreement between theoretical predictions
and measurements of the solar neutrino flux.
For many years, however, it was not possible to determine if the
observed discrepancy was due to problems with the experiment and/or
with the modeling of the Sun, or if it was, in fact, a sign of new physics.
In the last decade other neutrino experiments,
Kamiokande \cite{Kamiokande}, GALLEX \cite{GALLEX}, SAGE \cite{SAGE},
and more recently Super-Kamiokande \cite{Super-K},
have also measured the solar neutrino flux, with different
energy thresholds and using very different techniques. All four experiments
confirm a deficit in the observed number of solar neutrino induced
events. Moreover, it has recently become clear that it is virtually
impossible to
concoct a solar model which would fit all the data
\cite{bksreview,ssm_indep}.
On the other hand, the results of all experiments can be explained by
assuming that the electron neutrino oscillates into a different flavor state.
There are two neutrino oscillation scenarios that are capable of
faithfully explaining the solar neutrino data \cite{bksreview}.
One scenario makes use of
the MSW effect \cite{W,MS}, where the electron neutrino conversion into another
neutrino flavor is due to flavor dependent interactions with solar
matter. The other scenario assumes that the neutrino oscillation
length is comparable to the Earth-Sun distance, and simple
vacuum oscillations are sufficient. This scenario is also known as the
``just-so'' solution \cite{glashowkrauss}. Both scenarios allow the electron
neutrinos to oscillate into other active species or sterile neutrinos.
The solar neutrino energy spectrum is determined by several nuclear
reactions which take place in the Sun's core \cite{reactions}, and
different experiments are sensitive to neutrinos produced by different
nuclear reactions.
Super-Kamiokande, for example, a very large water Cherenkov
detector, is currently sensitive to solar neutrinos with energies
slightly above 5.5~MeV.
Almost all neutrinos it detects
come from the decay of $^8$B ($^8$B $\rightarrow$ $^8$Be$^{*} +
e^++\nu_e$).\footnote {There is a small fraction of the neutrinos
that can be detected at Super-Kamiokande coming from the reaction
$^3$He $+ p\rightarrow ^4$He $+e^+ + \nu_e$.}
Another solar reaction that gives rise to neutrinos is the process of
electron capture
by a $^7$Be nucleus ($^7$Be $+e^-\rightarrow$ $^7$Li
$+\nu_e$). Neutrinos from this reaction have energies below the
Super-Kamiokande threshold, but are accessible to the radiochemical
experiments Homestake, GALLEX, and SAGE. If one naively assumes that
the suppression in the neutrino flux is due to the suppression of
individual neutrino sources ($^8$B, $^7$Be, etc) in the Sun, the
combination of the Super-Kamiokande data with that of the radiochemical
experiments indicates that the flux of $^7$Be neutrinos is virtually
absent \cite{ssm_indep,noroom47Be} (the best fit value
of the $^7$Be flux is in fact negative!). In the case of neutrino
oscillations, all solutions to the solar neutrino puzzle indicate
that the $^7$Be neutrino flux is suppressed, in some cases very
strongly.
Thus, at present, there is great demand for
experiments that would accurately measure the flux of the $^7$Be
neutrinos. Two upcoming experiments, Borexino and KamLAND, may have
the capability to do exactly that.
In this paper, we present a quantitative study of what can be
accomplished by measuring the seasonal variations of the $^7$Be
neutrino flux at Borexino and KamLAND.
Seasonal variations of the solar neutrino flux are of course expected,
due to the Earth's eccentric orbit. The number of
neutrinos of all flavors reaching the Earth is larger when the Earth is
closer to the Sun than when it is farther away, and should vary as
$1/L^2$. In the case of no neutrino oscillations or of the MSW solution
to the solar neutrino puzzle, the number of $^7$Be solar neutrino
induced events is supposed to vary according to the $1/L^2$ law,
following the variation of the total neutrino flux.
This will be referred to as the ``normal'' seasonal
variation.
If vacuum oscillations are the solution to the solar
neutrino puzzle, large, anomalous
seasonal variations of the number of $^7$Be solar neutrino induced
events might be detected \cite{glashowkrauss,old_osc}.
It is well known that neutrino oscillation effects depend on
the distance to the neutrino source, and
different Earth-Sun distances may yield very different $\nu_e$
survival probabilities \cite{PP,sea_var}.
The anomalous seasonal variation effect should
be more pronounced in $^7$Be neutrinos than in $^8$B neutrinos (the
latter was recently studied in \cite{justsoback}). This
is due to one
important feature which distinguishes $^7$Be neutrinos
from $^8$B and other abundant types of solar neutrinos: because they are
produced as part of a two-body final state, the neutrino energy
spectrum is mono-energetic.\footnote{In fact there are two distinct
neutrino energies, 0.383 and 0.862 MeV, corresponding to different
final states of the $^7$Li nucleus. Borexino and KamLAND are only
sensitive to the higher energy component.} The details
will become clear when we discuss the anomalous seasonal variation
effect, in Sec.~\ref{sec:sensitivity}.
In the case of no anomalous seasonal variations,
if one has enough statistics and a small enough background, the
time variation of the data can be used to measure the solar
neutrino flux, given that the number of background events is
constant in time.\footnote{Actually, a time-dependent background is
also acceptable, as long as it can be monitored and understood well
enough.} We will analyze how well
Borexino and KamLAND can perform this type of measurement. We are
particularly interested in analyzing the relevance of this technique
when the number of electron neutrinos reaching the detector is very
suppressed with respect to the Standard Solar Model predictions, as
might be
the case if there are $\nu_e\rightarrow \nu_{\mu,\tau}$ oscillations
for the small angle MSW solution.\footnote{If $\nu_e$ oscillates into
sterile neutrinos, the suppression is even more pronounced, due to
the absence of neutral current $\nu_{\mu,\tau}$-$e$ elastic
scattering. We do not consider oscillations into sterile neutrinos
in this paper.}
The paper is organized as follows. In Sec.~\ref{sec:flux} we discuss how
seasonal variations might be used to determine the solar neutrino flux
at Borexino and KamLAND,
in such a way that no separate measurement of the number of background
events is required.
In Sec.~\ref{sec:sensitivity} we analyze the effect of the vacuum
oscillation solutions to
the solar neutrino puzzle on the annual variation of the number of
detected events at Borexino and KamLAND. In particular we describe the
region of the ($\sin^22\theta,\Delta m^2$) parameter space where
vacuum oscillations can be discovered by studying the seasonal
variations of the data.
In Secs.~\ref{sec:measurement} and \ref{sec:exclusion} we describe how
the measurement of the seasonal
variation of the $^7$Be solar neutrino flux may be used to either
measure the neutrino oscillation parameters, $\sin^22\theta$ and $\Delta
m^2$, or exclude a
large portion of the ($\sin^22\theta,\Delta m^2$) parameter
space. In Sec.~\ref{sec:conclusion} we summarize our results and conclude.
\setcounter{equation}{0}
\section{Measuring the $^7$Be Solar Neutrino Flux}
\label{sec:flux}
As was already pointed out, measuring the flux of $^7$Be neutrinos is
crucial towards understanding the solar neutrino puzzle.
Borexino \cite{Borexino}
plans to do this measurement by using 300 tons of organic liquid
scintillator to detect recoil electrons from elastic $\nu$-$e$
scattering. Since the scintillator has no directional
information, and the signal is characterized only by the scintillation
light produced by the recoil electron,
the background has to be kept under control. This places a very
stringent constraint
on the radio-purity of the scintillator and on the activity of
the material in the detector. Borexino anticipates 100 tons of fiducial
volume for detecting solar neutrinos.
KamLAND \cite{KamLAND}, which
was originally
conceived as a reactor neutrino experiment with an unprecedented
baseline (170~km on the average), may also be able to study $^7$Be solar neutrinos,
if rigorous yet attainable requirements on the radio-purity and activity
are met.
We assume throughout the paper that
KamLAND will use 600 tons of fiducial volume for detecting solar
neutrinos (the size of the fiducial volume will depend on the
background rate, which is currently unknown).
We concentrate our analysis on Borexino, which is an
approved dedicated solar neutrino experiment, and discuss KamLAND, whose
uses for solar neutrino studies are at present being proposed
\cite{kamlandprop},
as a possible higher statistics improvement.
It is important to define what is meant by ``measuring the $^7$Be
solar neutrino flux.''
In reality, what the experiments are capable of measuring is
the number of recoil electrons induced by solar neutrino
interactions in a given recoil electron kinetic energy range
(kinematic range).
This information can only be converted into a solar
neutrino flux measurement if one knows the flavor composition of the
solar neutrinos \cite{dG_M}. Explicitly, assuming that the
solar neutrino flux
is composed of $\nu_e$ (with fraction $P$) and $\nu_{\mu,\tau}$ (with
fraction $Q=1-P$),
\begin{equation}
\#{\rm recoil}\hspace{1mm}
{\rm electrons/time}
=\Phi\times(P\sigma_{\nu_e\mbox{-}e}+(1-P)\sigma_{\nu_{\mu,\tau}\mbox{-}e})N_e
,
\end{equation}
where $\Phi$ is the neutrino flux, $N_e$ is the number of target
electrons, and
\begin{equation}
\sigma_{\nu_{x}\mbox{-}e}\equiv\int_{T_{\rm
min}}^{T_{\rm max}}dT \left(\frac{{\rm d}\sigma}{{\rm
d}T}\right)_{\nu_{x}\mbox{-}e},
\end{equation}
with
$\left(\frac{{\rm d}\sigma}{{\rm d}T}\right)_{\nu_{x}\mbox{-}e}$ being
the
differential cross section for $\nu_x$-$e$ scattering for a given
kinetic energy $T$ of the recoil electron. $T_{\rm min}$ and $T_{\rm
max}$ define
the kinematic range.
In the case of neutrino oscillations, $P$ is the
the survival probability for electron neutrinos, while $1-P$ is the
probability that $\nu_{e}$ will oscillate into $\nu_{\mu,\tau}$.
If the flavor composition of the flux is not
known, all that can be quoted is the effective neutrino flux,
$\Phi_{\rm eff}$, which is calculated from the number of measured recoil
electrons assuming that there are only electron neutrinos coming from
the Sun. Explicitly,
\begin{equation}
\Phi_{\rm eff}\equiv\frac{\#\rm{recoil}\hspace{1mm}
\rm{electrons/time}}{\sigma_{\nu_e\mbox{-}e}N_e}=\Phi\times\left(
P+\left(1-P\right)\frac{\sigma_{\nu_{\mu,\tau}\mbox{-}e}}
{\sigma_{\nu_e\mbox{-}e}}\right).
\end{equation}
Clearly, if $P=1$, $\Phi_{\rm eff}=\Phi$. It is important to remember
that $\sigma_{\nu_{\mu,\tau}\mbox{-}e}/\sigma_{\nu_e\mbox{-}e}<1$
and therefore
$\Phi_{\rm eff}\leq\Phi$. The ratio of the neutrino elastic cross
sections depends on the energy of the incoming neutrino and the
kinematic range to which each particular experiment is sensitive.
For $E_\nu=0.862$~MeV and the Borexino (KamLAND) kinematic range
250--800~keV (280--800~keV),
$\sigma_{\nu_{\mu,\tau}\mbox{-}e}/\sigma_{\nu_e\mbox{-}e} = 0.213 (0.214)$.
It is this effective electron
neutrino flux, $\Phi_{\rm eff}$, that is referred to, throughout this
paper (and in general), as the $^7$Be solar neutrino flux.
In order to determine the number of recoil electrons induced by solar
neutrino interactions, it is crucial to determine
the number of background events. The number of background
events can be estimated by various techniques, which we do not address
in this paper. It is worthwhile to point out, however, that this is
a very difficult process and it would be highly desirable to have an independent
way to determine the $^7$Be solar neutrino flux in order to make the
final results more convincing. This may be possible if one looks at
the seasonal variation of the number of detected events.
In the following, we study the seasonal variation of the event rate as
a means to measure
the $^7$Be solar neutrino flux. The distance between the Earth and
the Sun varies slightly over seasons because of the eccentricity of
the Earth's orbit. The perihelion (when the Earth is closest to the
Sun) occurs around January first.
The eccentricity of
the Earth's orbit is $\epsilon = 0.017$, and hence the distance varies
as
\begin{equation}
L = L_0 ( 1 - \epsilon \cos (2 \pi t/{\rm year}) ),
\end{equation}
to the first order in $\epsilon$. Here, $t$ is the time measured in years
from the perihelion, and $L_0 = 1.496 \times 10^8$~km is one
astronomical unit. The neutrino flux varies as $1/L^2$ and hence
shows a seasonal variation of about 7\% from minimum to
maximum. The change in the Earth-Sun distance between the aphelion and
the perihelion is given by
\begin{equation}
\label{DeltaL}
\Delta L\equiv L_{{\rm max}}-L_{{\rm min}}=2\epsilon
L_0=5.1\times10^{6}\hspace{1mm}{\rm km}.
\end{equation}
By fitting the event rate to the seasonal variation expected
due to the eccentricity,
\begin{equation}
B + S \left(\frac{L_0}{L}\right)^2,
\end{equation}
one can extract the background event rate $B$ and the signal event rate $S$
independently. As long as the detector is monitored well and its
performance is sufficiently stable, this method will be only limited
by statistics.
Borexino expects 53
events/day\footnote{For simplicity, we neglect the contribution of solar
neutrino sources other than $^7$Be electron capture throughout the paper.
In particular we neglect the contribution of neutrinos produced
in the CNO cycle, which is about 10\% of that from the $^7$Be
neutrinos.}
according to the BP95 \cite{BP95} Standard Solar Model (SSM),
together with 19 background events/day \cite{Borexino}, after the statistical
subtraction of the known background sources.
This is done by pulse shape discrimination against the $\alpha$-particle
background and the measurement of Bi-Po pairs via $\alpha$-$\beta$
coincidence. This in turn allows the statistical subtraction of
processes in the
$^{238}$U and $^{232}$Th chains which are in equilibrium.
It is also assumed that the experiment can achieve a radio-purity of
$10^{-16}$g/g for U/Th, $10^{-18}$g/g for $^{40}$K, $^{14}{\rm
C}/^{12}{\rm C}= 10^{-18}$, and no Rn diffusion. For KamLAND we
use 466 events/kt/day for the signal and 217 events/kt/day
\cite{kamlandprop} for the
background under similar assumptions but with larger cosmogenic
background (especially $^{11}$C) and some Rn diffusion.
Assuming 600~t of fiducial volume, we expect 280 signal events/day and
130 background events/day.
Throughout the paper, we will assume that the number of background
events is either constant
in time or its time dependence is sufficiently well understood by
monitoring. We neglect systematic effects and assume that there are
only statistical uncertainties.
\begin{figure}[tp]
\centerline{
\psfig{file=seasonal-BOREXINO.eps,width=0.5\textwidth}
\psfig{file=seasonal-KamLAND.eps,width=0.5\textwidth}
}
\caption{The simulated seasonal variation of the $^7$Be flux for the case of the
small angle MSW solution, for three years of Borexino (left) and
KamLAND (right) running. The inset shows the measured flux of
$^7$Be neutrinos from the fit to the seasonal variation of the
event rate (point with error bar) and the SSM prediction (shaded band).}
\label{seasonalMC}
\end{figure}
Under these assumptions, Fig.~\ref{seasonalMC} depicts a simulation of
the seasonal variation of the ``data'' for both Borexino and KamLAND, after
three years of running. The plots
are for the case of the small angle MSW solution to the solar neutrino
puzzle, where the $\nu_e$'s
produced by $^7$Be electron capture inside the Sun
have almost completely oscillated into $\nu_\mu$ or $\nu_\tau$,
and the event rate is reduced to 21.3\% (21.4\%) of the SSM prediction
at Borexino (KamLAND).
In the fit to the data, both the
background and the $^7$Be flux are allowed to float.
This analysis can be repeated for different values of the $^7$Be flux, or,
equivalently, for different survival probabilities for
$\nu_e$. Fig.~\ref{flux-accuracy} depicts the expected 1~$\sigma$ statistical
accuracy of the $^7$Be flux measurement, together with the central value
normalized by the SSM prediction, as a function of the survival
probability for $\nu_e$. We emphasize that this measurement
technique assumes no knowledge of the background.
\begin{figure}[tp]
\centerline{
\psfig{file=7Be-BOREXINO.eps,width=0.5\textwidth}
\psfig{file=7Be-KamLAND.eps,width=0.5\textwidth}
}
\caption{The expected 1~$\sigma$ statistical accuracy of the $^7$Be neutrino flux
measurement, together with the central value normalized by the flux
predicted by the SSM, as a function of the $\nu_e$ survival
probability at Borexino (left) and KamLAND (right), after
three years of data taking.}
\label{flux-accuracy}
\end{figure}
The important information one should obtain from this analysis is if one can
indeed measure a nonzero $^7$Be solar neutrino flux.
For example, in the case of the small angle MSW solution, the $\nu_e$
survival probability is very close to zero and, assuming the expected
number of background events, Borexino's measured neutrino flux
is less than 1.5~$\sigma$ away from zero. The situation at KamLAND
is much better, and in the case of the small angle MSW solution a
healthy 3 sigma-away-from-zero measurement of the flux is obtained, if
the background is as low as expected. The significance of the measured
flux increases for larger survival probabilities, as in the case of
the large angle and the low $\Delta m^2$ MSW solutions.
\begin{figure}[tp]
\centerline{
\psfig{file=maxb.eps,width=0.5\textwidth}
\psfig{file=maxk.eps,width=0.5\textwidth}
}
\caption{The maximum number of background events allowed per day at Borexino
(left) or KamLAND (right), for 3 years of running, in
order to measure a solar neutrino flux which is 3~$\sigma$ away from
zero. The dashed lines indicate the currently anticipated number of
background events per day. }
\label{max_bkg}
\end{figure}
A similar analysis can be performed in order to determine how many
background events each experiment can tolerate in order to claim
a solar neutrino flux measurement which is 3~$\sigma$ away from
zero. Fig.~\ref{max_bkg} depicts the maximum number of background
events per day allowed for 3 years of Borexino or KamLAND
running. It is worthwhile to comment that, in the case of Borexino and the
small angle MSW solution ($P\simeq0$), a 3 sigma-away-from-zero measurement of the
neutrino flux is not attainable
in three years, even in the case of no background (note that for
$P\lesssim0.05$ the required maximum background to achieve a three
$\sigma$ measurement of the flux is negative, {\it i.e.}, impossible to achieve).
Therefore, for
Borexino, this simple, background independent analysis using the
seasonal variation of the data is not particularly powerful in the
case of the small angle MSW solution,
due to statistical limitations.
\setcounter{equation}{0}
\setcounter{footnote}{0}
\section{Sensitivity to Vacuum Oscillations}
\label{sec:sensitivity}
In this section we study the discovery potential of the Borexino and
KamLAND experiments in the region of $\Delta m^2$ corresponding to the
vacuum oscillation solution to the solar neutrino problem.
In this case, the pattern of seasonal variations can be very distinct
from the normal pattern discussed in the previous section.
The basic idea is the following.
The survival probability $P$ for an electron neutrino in the case of
neutrino vacuum oscillations between two flavor states\footnote{One can
assume the more complicated case of oscillations between three
neutrino flavor states. In this paper we limit our studies to the
case of oscillations between two flavor states.} is given by
\begin{equation}
\label{osc}
P = 1 - \sin^2 2\theta \sin^2 \left(1.27 \Delta m^2 \frac{L}{E}\right),
\end{equation}
where the neutrino energy $E$ is in GeV, the distance $L$ in km, and the
difference of masses-squared $\Delta m^2$ in eV$^2$.
Model-independent analyses of all solar neutrino data show the
need for an energy-dependent suppression of the $\nu_e$ flux.
The ``just-so'' solution achieves this by choosing $\Delta
m^2$ such that the corresponding neutrino oscillation length
\begin{equation}
\label{eq:osclength}
L_{\rm osc}\equiv \frac{\pi E}{1.27\Delta m^2}=
2.47 \times 10^8 \mbox{ km} \times
\left(\frac{E}{10 \mbox{ MeV}}\right)
\left(\frac{10^{-10} \mbox{ eV}^2}{\Delta m^2}\right)
\end{equation}
is of the order of one Astronomical Unit (1~a.u. = $1.496 \times
10^8$ km); hence the name ``just-so''. More specifically, the
oscillation length is assumed comparable to 1~a.u. for $^{8}$B
neutrinos ($E_{\nu}\approx 10$~MeV); at the same time, the
oscillation length of $^7$Be neutrinos ($E_{\nu} = 0.862$~MeV) is
an order of magnitude smaller and, for sufficiently large $\Delta
m^2$, can be comparable to the seasonal variation of the Earth--Sun
distance due to the eccentricity of the Earth's orbit, $\Delta L$ (see
Eq.~(\ref{DeltaL})). As a
consequence, the flux of $^7$Be neutrinos detected on the Earth may
exhibit an anomalous seasonal variation, beyond the normal $1/L^2$
effect discussed in the previous section.
Such anomalous variation could serve as a unique signature of
vacuum oscillations \cite{glashowkrauss, old_osc}.
Moreover, as we will show in this section,
both Borexino and KamLAND will be able to cover a large portion of
the ``just-so'' parameter space, even without relying on a
particular solar model or estimate of the background rate, just by
analyzing the {\it shape} of their data.
In this sense the discovery of an anomalous seasonal variation at one
of these experiments would be as robust a result as the
Super-Kamiokande measurement of the up-down asymmetry for the
atmospheric muon neutrinos.
\begin{figure}[t]
\centerline{
\psfig{file=fig2_05.eps,width=1\textwidth}}
\caption{ Illustration of the effect of
vacuum oscillations on the shape of the seasonal
variation of the solar neutrino data.
The points with statistical error bars represent the number of
events/month expected at Borexino after 3 years of running for
$\Delta m^2 = 3 \times 10^{-10}$~eV$^2$, $\sin ^2 2\theta =
1$.
The histogram in (a) shows the number of events predicted
by the SSM without neutrino oscillations, plus the number of
anticipated background events.
The histogram in (b) shows the same quantity
after adjusting the solar neutrino flux
and the background rate so as to
minimize the value of $\chi^2$, as explained in the
text. The difference between the case with oscillations
and the one without oscillations is still apparent. }
\label{fig:illustratepoint}
\end{figure}
To illustrate the main idea, we choose a particular point
($\Delta m^2 = 3 \times 10^{-10}$~eV$^2$, $\sin ^2
2\theta = 1$) in the allowed region of the ``just-so'' parameter
space\footnote{Based on the analysis of the total rates in the
Homestake, GALLEX, SAGE, and Super-Kamiokande experiments. See
Fig.~5 in \cite{bksreview}.} and compute the corresponding seasonal
distribution of the neutrino events at Borexino after 3 years of
running. We use the number of background events and the expected number of
signal events (before the effect of neutrino oscillations) quoted in
Sec.~\ref{sec:flux}.
The results are shown in Fig.~\ref{fig:illustratepoint} by
the set of ``data'' points with error bars; each point represents
the number of events expected in a given month and the vertical error
bars show the corresponding statistical uncertainties. The histogram in
Fig.~\ref{fig:illustratepoint}(a) shows ``theoretical'' event rates
expected for non-oscillating neutrinos, provided the background rate
is known accurately and the SSM prediction for the neutrino flux is
trusted. One can see that under these assumptions vacuum neutrino
oscillations with $\Delta m^2 = 3 \times 10^{-10}$~eV$^2$,
$\sin ^2 2\theta = 1$ would be trivial to discover.
More importantly, the experiment would be able to claim the
discovery even without relying on an estimate of the background rate
or the value of incoming neutrino flux predicted by the SSM.
It is intuitively obvious from the figure that the
vacuum oscillation ``data'' points cannot be fit by the
``theoretical'' curve even if the background and the solar neutrino
flux are varied freely, unless one assumes neutrino
oscillations. This can be quantified as follows. For a given
background rate $b$ and signal event rate $s$, we define the
$\chi^2$ value of the fit for an ``average'' experiment:
\begin{equation}
\label{chi2}
\chi^2(s,b) = N_{\rm{d.o.f.}} + \sum_i^{N_{\rm{bins}}} \frac{(d_i - b - s\cdot h_i)^2}{d_i},
\end{equation}
where $N_{\rm{bins}}$ is the number of bins, $N_{\rm{d.o.f.}}$ is the number of
degrees of
freedom, $d_i$ is the {\it average} expected number of neutrino
events in the $i$th bin, and $h_i$ is given by $ h_i = \int_{i-1}^i ( 1 -
\epsilon \cos (2 \pi x/N_{\rm{bins}}) )^2 dx$ .
The constant term $N_{\rm{d.o.f.}}$ in Eq.~(\ref{chi2}) is added
to take into account the effect of statistical fluctuations in the
data.
In a single experiment, statistical
fluctuations make the number of neutrino events in the $i$th
bin slightly different from $d_i$, and $\chi^2$ is computed by an
expression similar to Eq.~(\ref{chi2}), with $d_i$ replaced by the number of
events measured in the $i$th bin and {\it without} the constant term,
$N_{\rm d.o.f.}$.
In our analysis, however, we are
interested in the sensitivity of an ``average'' experiment. As
proven in Appendix \ref{explainchi2}, averaging over many
experiments results in the definition of $\chi^2$ given in
Eq.~(\ref{chi2}), with the constant term $N_{\rm{d.o.f.}}$. This agrees
with the conventional wisdom that, if a function describes data
correctly, the average expected value of $\chi^2$ should be equal
to the number of degrees of freedom. Given this definition, we can
choose values of $s$ and $b$ that minimize the $\chi^2$; the only
restriction imposed is that both $s$ and $b$ be non-negative. For the case
at hand the minimum occurs when $b$ is zero and $s$ is
0.95 times the SSM prediction (see Fig.~\ref{fig:illustratepoint}(b)).
As expected, even after this change
the ``data'' points and the histogram are very different.
(Numerically, $\chi^2=2935$ which for 10 degrees of freedom implies
a confidence level of $1-9 \times 10^{-626}$!).\footnote{This number is,
of course, unrealistic, and the true confidence level in this case will be
dominated by systematic effects.}
We now extend this approach, and scan the entire ($\sin ^2 2\theta$,
$\Delta m^2$) plane
(for an earlier work with a more simplified analysis which does
not consider the presence of background, see \cite{PP}).
In the analysis below, we follow the same steps
as before: the ``data'' is simulated according to the expected number
of background and signal events, plus the effect of
neutrino oscillations, for each value of ($\sin ^2 2\theta$, $\Delta
m^2$), binned into a certain
number of bins $N_{\rm{bins}}$, and then compared to the
``theoretical'' predictions for the case of no oscillations. The
$\chi^2$ is computed according to Eq.~(\ref{chi2}) and minimized with
respect to both the signal ($s$) and background ($b$). The confidence
level (CL) corresponding to the
minimal value of $\chi^2$ and $N_{\rm d.o.f.}=N_{\rm{bins}}-2$
degrees of freedom is then determined, and the region in which the CL
is less than a given number is isolated. This case, when both
the number of background events and the incoming solar neutrino flux
are considered unknown in the fit, is the most conservative one, and yields the
smallest sensitivity region. Later we also study
less conservative cases, where we assume in the ``data'' analysis that the
incoming neutrino flux is the one predicted by the SSM and/or that the
background rate is known.
\begin{figure}[t]
\centerline{
\psfig{file=fig2_1.eps,width=0.7\textwidth}}
\caption{The sensitivity region of the Borexino experiment
in 3 years, if the analysis does not assume any knowledge
of the background rate or the incoming solar neutrino flux.
In the unshaded region
the ``data'' is at least 5~$\sigma$ away from the best
no-oscillations fit. In the lightly shaded region the
discrepancy is greater than $95\%$~CL but less than 5~$\sigma$~CL.}
\label{fig:lowerBorex}
\end{figure}
We now apply this most conservative procedure to study the
experimental reach of Borexino after 3 years of operation.
In Figure \ref{fig:lowerBorex} we show the results of the scan for
$95\%$ and 5~$\sigma$ CL. As one can see from the
figure, even at 5~$\sigma$ CL a large portion of the parameter space
above $\Delta m^2 \sim 10^{-10}$~eV$^2$ is covered (white region). In this
region the neutrino oscillation length $L_{\rm{osc}}$ is smaller than
the seasonal variation of the Earth--Sun distance $\Delta L$.
On the other hand, below $\Delta m^2 \sim
10^{-10}$~eV$^2$ one can see a series of spikes protruding through the
sensitivity region. It is important to understand the origin of these
spikes.
Since we adjust the level of signal and background in the fit,
we are not sensitive to the absolute event rate,
only to its variation during the year. For $\Delta m^2 \mathrel{\mathpalette\vereq<}
10^{-10}$~eV$^2$ the oscillation length is larger than $\Delta L$ and the
amplitude of the variation of the event rate is roughly proportional
to the first derivative of Eq.~(\ref{osc}) with respect to $L$. In the
regions where this derivative nearly vanishes, the amplitude of the
variations is small and the signal is indistinguishable from the case
of no oscillations.
This explains why the loss of sensitivity occurs not only when the
neutrinos undergo approximately an integer number of oscillations as
they travel to the Earth ($\Delta m^2=n\times 0.143\times
10^{-10}$~eV$^2$),
but also when the number of oscillations is close to a
{\it half-integer} ($\Delta m^2=(n+1/2)\times 0.143\times
10^{-10}$~eV$^2$).
In the latter case the absolute neutrino
flux is maximally suppressed, but the magnitude of the seasonal
variation is small.\footnote{Notice that the regions preferred from
the global fits have the absolute $^7$Be neutrino flux suppressed. See
Figs. \ref{fig:finalBorexino} and \ref{fig:finalKamLAND}.}
Given this explanation, one would expect that the spikes corresponding
to a half-integer number of oscillations should become shorter if in the
analysis we choose to rely on the SSM prediction of the incoming
neutrino flux and/or on the anticipated background rate.
It is straightforward to incorporate the knowledge of both
quantities and their uncertainties in our procedure. For
example, to impose the value of the incoming neutrino flux predicted
by the SSM, we
modify the expression of $\chi^2$ in Eq.~(\ref{chi2}) by adding
an extra term:
\begin{equation}
\label{modchi2}
\chi^2(s,b) \longrightarrow \chi^2(s,b) + \frac{(s-s_0)^2}{\sigma^2_{s_0}} ,
\end{equation}
where $s$ and $b$ are the values of the signal and background with respect
to which we later minimize $\chi^2$, $s_0$ is the SSM prediction for the
signal, and $\sigma_{s_0}$ is the uncertainty in $s_0$. The rest of
the analysis is carried out unchanged, except that the number of degrees of
freedom is increased by one to $N_{\rm{d.o.f.}}=
N_{\rm{bins}} - 1$. To use both the incoming flux
predicted by the SSM and the anticipated background rate,
two terms are added to
Eq.~(\ref{chi2}) and the number of degrees of freedom is increased
by two to $N_{\rm{d.o.f.}}=N_{\rm{bins}}$.
\begin{figure}[t]
\centerline{
\psfig{file=fig2_2.eps,width=0.7\textwidth}}
\caption{The sensitivity reach of the Borexino experiment
after 3 years of running (at 95$\%$ confidence level). The
three cases considered are: no knowledge of either
the background rate or the incoming solar neutrino flux
(the covered region is white); assumption that the incoming
solar neutrino flux is the one predicted by the SSM, with
9$\%$ uncertainty (the covered
region is white + light gray); assumption that the
background rate is known with 10\% uncertainty and the
incoming neutrino flux agrees with the SSM, with 9\%
uncertainty (the covered region is white + light
gray + medium gray).}
\label{fig:lowerBorexssmbg}
\end{figure}
The results of the calculation are shown in Fig.~\ref{fig:lowerBorexssmbg}.
The uncertainty on the solar model prediction of the
$^7$Be neutrino flux is taken to be $9\% $ \cite{bp98}, while the
uncertainty on the background is $10\%$ \cite{kamlandprop}.
As expected, the odd--numbered spikes do become shorter. The one
possibility not shown in the plot is the situation when one only assumes
knowledge of the background rate. In this case the spikes become
significantly thinner, although their length remains virtually unchanged.
In order to extend this analysis to values of $\Delta
m^2 > 10^{-9}$~eV$^2$, several issues must be confronted.
We will next address these issues one by one, and illustrate
the discussion in Fig.~\ref{fig:lnwdthandcore}.
\begin{figure}[t]
\centerline{
\psfig{file=fig2_3final.eps,width=0.7\textwidth}}
\caption{The relative roles of the binning effect, the
linewidth effect, and the matter effect, as
explained in the text. }
\label{fig:lnwdthandcore}
\end{figure}
The first and the most obvious point is that the number of bins needs to be
changed. The reason is that the frequency of the seasonal variations
increases with $\Delta m^2$, and above some value ($\Delta
m^2\simeq 8 \times
10^{-10}$~eV$^2$, for 12 bins) integration over the bin size washes out
the effect. To avoid this, we change the number of bins from 12 to
365. After the change, the effect of binning
kicks in at $\Delta m^2\simeq 2.4 \times 10^{-8}$~eV$^2$, as curve 1 in
Fig.~\ref{fig:lnwdthandcore} illustrates.
Next, there are two physical effects one must take
into account: one is the interaction of the neutrinos with solar
matter (the MSW effect), and the other is
the finite width of the $^7$Be solar neutrino line. One may worry
about the wash-out of the seasonal variation effect due to the finite
size of Sun's core. However, matter effects
make the core size effect irrelevant because
the mixing angle in the Sun's core is small and the oscillations
effectively start at the level-crossing point (see
Eq.~(\ref{eq:arrival})).\footnote{We thank E.~Lisi and L.~Wolfenstein
for pointing this out to us. For earlier papers on this particular
point, see \cite{W}, \cite{glashowkrauss}, and in particular,
\cite{Pantaleone}.}
When a $\nu_e$ is created by the electron capture process
in the core of the Sun, its Hamiltonian is dominated by the matter
effect $\sqrt{2} G_F n_e$ ($n_e$ is the electron number density) if
$\Delta m^2 \ll 10^{-5}~{\rm eV}^2$ for $^7$Be neutrinos.
We restrict ourselves to $\Delta m^2 < 10^{-7}$~eV$^2$ in the
following discussions, as the final sensitivity due to the anomalous
seasonal variation is limited by $\mathrel{\mathpalette\vereq<}
10^{-8}$~eV$^2$ as will be seen later in this section. Then the mass
mixing effect can
be completely ignored at the time of the neutrino production, and one
can safely take the produced neutrino to be in a Hamiltonian
eigenstate (the one which corresponds to the larger energy in the
Sun's core).
As it propagates through the Sun, the neutrino follows the instantaneous
Hamiltonian eigenstate (in the adiabatic approximation), and exits in
the heavier mass eigenstate, $\nu_2 = \nu_e \sin \theta + \nu_\mu \cos
\theta$. It also has a finite amplitude $A_c$ for hopping to the other
Hamiltonian eigenstate.
The neutrino state that exits the Sun can therefore be written as
\begin{equation}
\nu_{\rm exit} = A_c \nu_1 + B_c \nu_2,
\end{equation}
with the unitarity constraint $|A_c|^2 + |B_c|^2 = 1$. Out of the
Sun, the two mass eigenstates develop different phases due to the
mass difference, $e^{-i \Delta m^2 t / 2 E_\nu}$. Therefore the
neutrino state that arrives at the Earth is given by
\begin{equation}
\nu_{\rm arrival} = A_c \nu_1 + B_c \nu_2 e^{-i \Delta m^2 L/2E_\nu},
\label{eq:arrival}
\end{equation}
up to an overall phase factor.
The distance $L$ is between the point of level crossing and the
Earth. Finally, the survival probability of the electron neutrino is
determined by the $\nu_e$ component of $\nu_{\rm arrival}$, and
hence
\begin{eqnarray}
P &=& |A_c \cos \theta + B_c \sin \theta e^{-i \Delta m^2 L/2E_\nu}|^2
\nonumber \\
& = & |A_c|^2 \cos^2 \theta + |B_c|^2 \sin^2 \theta
+ 2 {\rm Re} A_c^* B_c e^{-i \Delta m^2 L/2E_\nu} \sin\theta \cos\theta .
\end{eqnarray}
Since $|B_c|^2$ is the hopping probability between two Hamiltonian
eigenstates in the Sun $P_c$, one can rewrite the formula using $P_c$
and an additional phase factor $A_c^* B_c = \sqrt{P_c (1-P_c)}
e^{-i\delta}$,
\begin{equation}
P = P_c \cos^2 \theta + (1-P_c) \sin^2 \theta
+ 2 \sqrt{P_c (1-P_c)} \sin\theta \cos \theta
\cos \left( \frac{\Delta m^2 L}{2E_\nu} + \delta \right).
\label{eq:withMSW}
\end{equation}
An approximate formula for $P_c$ was given in \cite{Petcov} using the
exponential density profile of the Sun,
\begin{equation}
P_c = \frac{e^{-\gamma \sin^2 \theta}-e^{-\gamma}}{1-e^{-\gamma}}
\label{eq:Petcov}
\end{equation}
with
\begin{equation}
\gamma = 2\pi r_0\frac{\Delta m^2}{2 E_{\nu}}=1.22\left(\frac{\Delta
m^2}{10^{-9}{\rm eV}^2}\right)\left(\frac{0.862{\rm MeV}}{E_{\nu}}\right),
\end{equation}
where we consider the exponential-profile approximation for the
electron number density in the Sun $n_e \propto \exp(-r/r_{0})$, with
$r_0=R_\odot/10.54=6.60\times 10^{4}$~km, given in \cite{Bahcallbook}.
Fig.~\ref{fig:Pc} shows the contours of $P_c$ on the $(\sin^2 2\theta,
\Delta m^2)$ plane for the $^7$Be neutrino energy $E_\nu = 0.862$~MeV.
\begin{figure}
\centerline{
\psfig{file=Pc.eps,angle=90,width=0.7\textwidth}}
\caption{The contour plot of the hopping probability $P_c = 0.1$,
0.2, $\ldots$, 0.9, for the $^7$Be neutrino energy, using the
exponential-profile approximation for the electron number density
and Eq.~(\ref{eq:Petcov}).}
\label{fig:Pc}
\end{figure}
The most important consequence of the matter effect is that the vacuum
oscillation is suppressed when $P_c \rightarrow 0$ (adiabatic limit).
The origin of the suppression is simple. When $P_c$ is small, the
neutrino state that exits the Sun is nearly a pure $\nu_2$ state. Since
it is a mass eigenstate, only its phase evolves in time and no
oscillations take place. The $\nu_e$ survival probability then is
simply given by the $\nu_e$ content of $\nu_2$, which is nothing but
$\sin^2 \theta$, without anomalous seasonal variations. Therefore, the
sensitivity to the anomalous seasonal variation is reduced in the region
with small $P_c$. When $\Delta m^2$ is small, on the other hand, the situation
is in the extreme non-adiabatic limit, and $P_c \rightarrow \cos^2
\theta$. Then Eq.~(\ref{eq:withMSW}) reduces
to Eq.~(\ref{osc}). As $\Delta m^2$ increases, $P_c$ becomes smaller
than $\cos^2 \theta$, which enhances the vacuum oscillation effect in
the small mixing angle region. Curve 2 in
Fig.~\ref{fig:lnwdthandcore} includes the matter effect and indeed
indicates a reduced sensitivity for large $\sin^2 2\theta$ (small $P_c$)
and an enhanced sensitivity for small $\sin^2 2\theta$ (where $P_c$
starts deviating from $\cos^2 \theta$).\footnote{In the numerical
scan, we ignored the additional phase factor $\delta$, because its
effects are negligible \cite{Pantaleone}.}
The second effect is the finite width of the $^7$Be line. To give some
preliminary
idea about the relative size of this effect, we first consider a simplified
model. We assume for a moment that the only source of the line
broadening is the Doppler shift of
neutrino energies arising from the thermal motion of the $^7$Be
nuclei. Since the energy is shifted to $E\rightarrow E(1+v_z/c)$
and the probability distribution of the velocity along the line of
sight $v_z$ is proportional to $\exp(-m
v_z^2/2kT)$, the resulting line profile will be a Gaussian $\exp(-m
c^2 (E-E_0)^2/(2kT E_0^2))$.
Taking the temperature to be 15.6 million Kelvin (the temperature
in the center of the Sun) and integrating over the
line profile, we obtain curve 3 in Fig.~\ref{fig:lnwdthandcore}. The sensitivity loss
now occurs at $\Delta m^2\simeq 1 \times 10^{-8}$~eV$^2$,
demonstrating that this effect is more
important than the matter effect.
This naive model is actually incomplete; there exists another very
important source of line broadening. Because the incoming electron in
the process $^7\mbox{Be} + e^- \rightarrow {^7\mbox{Li}} + \nu_e$
has nonzero thermal kinetic energy, the center of mass energy of
the reaction is greater than the one measured in the laboratory, and so the
neutrino has a greater energy.
The phase space distribution of electrons is governed by the
Maxwellian factor $\exp(-E_{e^-}/k T)$. This distribution has to be
multiplied by the energy-dependent cross section, integrated over the
phase space, and finally convoluted with the Gaussian arising from the
Doppler effect.
The resulting line
shape becomes asymmetric, with a Gaussian profile on the left (due to
the Doppler effect) and an exponential tail on the right (due to the
Maxwellian distribution of the electron energy). The issue was studied
in detail in \cite{lineprfl}, where the precise form of the profile was
computed.\footnote{It turns out that other effects, such as
collisional line broadening \cite{loeb} or gravitational
energy shift \cite{lineprfl}, are unimportant.}
Repeating the calculation with this profile we generate curve 4 in
Fig.~\ref{fig:lnwdthandcore}.
One can see that for this curve the cut-off
occurs at smaller $\Delta m^2$. This behavior
is expected, because the linewidth is now greater than when only the
Doppler effect was included (curve 3 in
Fig.~\ref{fig:lnwdthandcore}).
It is also worth noting that the cut-off
sets in more gradually. This feature can be understood
analytically by considering the Fourier transform of an exponential tail
vs. a Gaussian tail. The details can be found in Appendix \ref{app2}.
Finally, we can combine both the linewidth and the matter
effects. The result is curve 5 in Fig.~\ref{fig:lnwdthandcore}.
As expected, the inclusion of the matter effect on top of the
linewidth effect introduces only a small distortion to the sensitivity
region. It is important to
note that for $\Delta m^2\mathrel{\mathpalette\vereq<} 5\times 10^{-10}$~eV$^2$ none of the
physical effects mentioned above affect the sensitivity region (curve
1 versus curve 5, in Fig.~\ref{fig:lnwdthandcore}).
We need to consider one last ingredient in the analysis. We
again return to the issue of the number of bins. While choosing more
bins is necessary for larger values of $\Delta m^2$, it simultaneously
leads to a loss of sensitivity for smaller $\Delta m^2$.
A better procedure is to use an optimum number of bins
$N_{\it opt}$ for each $\Delta m^2$. It can be shown that for our method of
analysis (minimizing $\chi^2$ by varying the signal and background)
and sufficiently large $\Delta m^2$ an
approximate formula holds: $N_{\it opt} \simeq 2 \times 10^{10} (\Delta m^2
/1\mbox{ eV}^2)$. Of course, this formula should not be used when
the optimal number of bins it predicts is too small.
We choose to use 12 bins for
$\Delta m^2 \leq 6 \times 10^{-10}$~eV$^2$ and a variable
number of bins $N_{\rm{bins}} = 2 \times 10^{10} (\Delta m^2
/1\mbox{ eV}^2)$ for $\Delta m^2 > 6 \times
10^{-10}$~eV$^2$.\footnote{An alternative technique, which can be
considered more rigorous but
which would also be more computer intensive, is to Fourier transform
the simulated data for every value of ($\sin^22\theta,\Delta m^2$) in
the scan. One can then compare the intensities of the harmonics to
those expected for the case of no oscillations. A
description of this method can be found in \cite{fourier}. For our
purposes varying the number of bins is sufficient.}
In Fig.~\ref{fig:finalBorexino} we show the entire sensitivity reach
of Borexino after three years of running.
The unshaded region will be covered at
least at 95$\%$ CL, if in the analysis one allows the background and
the incoming solar neutrino flux to float. The
dark shading marks the additional portion of the parameter space that
will be covered at least at 95$\%$ CL, if in the analysis one assumes both the
anticipated background rate (10$\%$ uncertainty) and the SSM
prediction of the $^7$Be solar neutrino flux (9$\%$ uncertainty).
For $\Delta m^2 \mathrel{\mathpalette\vereq>} 5 \times 10^{-9}~{\rm eV}^2$, the sensitivity
to the anomalous seasonal variation gets lost because of the
smearing due to the linewidth effect. However, there is an overall
suppression of the flux due to the MSW effect in this region.
To be sensitive to this overall suppression, we should return
to a smaller number of bins to enhance the statistical accuracy. We
therefore use 12 bins in this region.\footnote{One can cover a
slightly larger portion of the parameter space by using yet fewer
bins. We chose 12 bins such that one can still verify the expected
$1/L^2$ behavior of the signal even with a reduced flux, as we
discussed in Sec.~\ref{sec:flux}.}
For comparison, we also superimpose
the ``just-so'' preferred regions obtained by analyzing the total event
rates in the Homestake, GALLEX, SAGE, and Super-Kamiokande
experiments (Fig.~5 in \cite{bksreview}). The plot shows
that Borexino will be sensitive to almost all of the preferred
region, even without relying on the SSM prediction of the incoming
neutrino flux or on the knowledge of the background rate. Only two thin spikes
protrude through the lower ``islands''. This overlap disappears
completely when the anticipated background rate and the SSM prediction
for the incoming neutrino flux are used in the ``data'' analysis,
in which case the entire preferred region is covered.
\begin{figure}[tp]
\centerline{
\psfig{file=fig2_4_1final.eps,width=0.9\textwidth}}
\caption{The final sensitivity plot for three years of
Borexino running, after
the inclusion of all effects limiting the reach of the
experiment for large $\Delta m^2$. The white region corresponds
to the sensitivity at more than 95$\%$ confidence level
with both the incoming neutrino flux and background rate
assumed to be unknown, and the dark region to the
additional coverage when the SSM $^7$Be flux and the
background rate estimated elsewhere are used. Also shown are
the regions preferred by the analysis of the total
rates in the Homestake, GALLEX, SAGE, and Super-Kamiokande
experiments \cite{bksreview}.}
\label{fig:finalBorexino}
\end{figure}
Fig.~\ref{fig:finalKamLAND} contains a similar plot for three years of
KamLAND running. Because KamLAND will have more statistics, it will be
sensitive at
$95\%$ CL to the entire preferred region without relying in the
analysis on the SSM prediction of the incoming neutrino flux or on
the knowledge of the background rate.
\begin{figure}[tp]
\centerline{
\psfig{file=fig2_5_1final.eps,width=0.9\textwidth}}
\caption{The same as Fig \ref{fig:finalBorexino}, but for
three years of KamLAND running.}
\label{fig:finalKamLAND}
\end{figure}
As mentioned earlier, the
sensitivity to anomalous seasonal variations is completely lost for
$\Delta m^2\mathrel{\mathpalette\vereq>} 10^{-8}$~eV$^2$. In this case
the seasonal variation of the data is consistent with an average
suppression of the incoming neutrino flux. In particular, in the case
of the MSW solutions ($10^{-7}\mbox{ eV}^2\mathrel{\mathpalette\vereq<}\Delta m^2\mathrel{\mathpalette\vereq<}
10^{-4}\mbox{ eV}^2$), no anomalous seasonal variations can be
detected, as was implicitly assumed in Sec.~\ref{sec:flux}.
At last, it is worth mentioning that the experiments will still be
sensitive to a significant part of the preferred region even if the
background rate or the incoming $^7$Be neutrino flux (for all flavors)
turns out to be significantly different.
For example, if the background rate at Borexino (KamLAND) turns out to be
30 (100) times higher than expected, the part of the preferred region with
$\Delta m^2 > 10^{-10}$~eV$^2$ will still be within the reach of the
experiment, after three years of running.
The sensitivity will be completely lost only if the background rate turns
out to be three (four) orders of magnitude higher than anticipated at
Borexino (KamLAND).
The consequences of a $^7$Be solar
neutrino flux smaller than predicted by the SSM can also be studied.
If the $^7$Be neutrino flux is for some reason
suppressed by a factor of 5, KamLAND is still sensitive to the part of
the preferred region with $\Delta m^2 > 10^{-10}$~eV$^2$, after 3
years of running.
\setcounter{equation}{0}
\section{Measuring the Oscillation Parameters}
\label{sec:measurement}
In this section, we address the issue of how well the two-neutrino
oscillation parameters, $\sin^22\theta$ and $\Delta m^2$, can be
extracted if the data collected at future solar neutrino experiments
exhibits an anomalous seasonal variation.
In order to do this, we simulate ``data'', according to the procedure
developed in Sec.~\ref{sec:sensitivity}, for two distinct points in
the parameter
space, $\sin^22\theta=0.7$, $\Delta m^2=8\times
10^{-11}$~eV$^2$ (``low point'') and $\sin^22\theta=0.9$,
$\Delta m^2=4.5\times 10^{-10}$~eV$^2$ (``high point'').
The low point is close to the best
fit point presented in \cite{bksreview}, while the high point is close
to the point preferred by the Super-Kamiokande analysis of the recoil
electron energy
spectrum \cite{Super-K_spectrum}. The data is binned into months (12
bins per years), and Fig.~\ref{points} depicts the annual variations
for both the high and the low points, assuming three years of Borexino
running. The no-oscillation case is also
shown.
\begin{figure}[t]
\centerline{
\psfig{file=year.eps,width=0.6\textwidth}
}
\caption{Number of recoil electrons detected in a given
month, for the low point, the high point (see text for description)
and the case of no neutrino oscillations, after three years of
Borexino running.}
\label{points}
\end{figure}
In order to measure the oscillation parameters, we
perform a 4 parameter ($s$, $b$, $\sin^22\theta$, and $\Delta m^2$)
fit to the ``data''. The fit is performed by minimizing $\chi^2$
with respect to the incoming neutrino flux ($s$) and the background
rate ($b$), as in Sec.~\ref{sec:sensitivity}, and computing it for
fixed $\sin^22\theta$ and
$\Delta m^2$. Fig.~\ref{measurement} depicts the
values of ($\sin^22\theta,\Delta m^2$) and the 95\% CL contours
(for two degrees of freedom), extracted from the ``data''
consistent with the low (light) and high (dark) points. Note that this
is very different from what was done in the previous section. There,
for each point in the ($\sin^22\theta$,$\Delta m^2$) plane there was a
different ``data'' set, and the ``data'' was fitted by a
non-oscillation theoretical function. Here the ``data'' is fixed
(either the low or the high point), and
is fitted by a theoretical function which assumes neutrino
oscillations.
\begin{figure}[t]
\centerline{
\psfig{file=measure2.eps,width=0.7\textwidth}
}
\caption{Measurement of the neutrino oscillations parameters
$\sin^22\theta$ and $\Delta m^2$,
assuming no knowledge of the SSM and the number of background
events. The regions represent the 95\% confidence level contours,
for data consistent with the high (dark) and low points
(light). The input points are indicated in the figure by the two crosses.
See text for details. We assume 3 years of Borexino running.}
\label{measurement}
\end{figure}
One should easily note that the extracted 95\% CL
contour for the high point consists of only two ``islands'', while for the
low point one extracts a collection of ``islands''. The reason for this is
simple. When $\Delta m^2\sim \rm{few}\times10^{-10}$~eV$^2$, the
oscillation length is slightly smaller than
$\Delta L$ (see Eq.~(\ref{DeltaL})).
This means that the seasonal variation of the ``data'' has a very
particular shape (as one may easily confirm by looking at
Figs.~\ref{fig:illustratepoint}, \ref{points}), which cannot be easily
mimicked by other values of $\Delta m^2$, even when the background
rate and the incoming flux are varied in the fit procedure.
When $\Delta m^2\sim \rm{several}\times10^{-11}$~eV$^2$, the oscillation
length is larger than $\Delta L$, and the
effect of seasonal variations is less pronounced. There is
a collection of $\Delta m^2$'s that yields the same
qualitative behavior. Because our fit procedure allows for the
background rate and the neutrino flux to float freely, a good agreement
with the ``data'' is met for a large portion of the parameter space.
In order to make this discussion clearer, it is useful to describe
in detail what happens to the number of electron neutrinos reaching
the detector as a function of time.
In the case of the low point: initially, when the Earth is at the
perihelion, the $\nu_e$ survival probability is small and, as time
progresses, monotonically increases until the Earth reaches the
aphelion (after six
months). The process happens in reverse order in the next six months, as
expected.
There are many other values of the oscillation length, {\it
i.e.} $\Delta m^2$,
such that the survival probability monotonically increases for
increasing Earth-Sun
distance and therefore a similar qualitative behavior is to be
expected. The main quantitative difference is in the
ratio of the number of events detected in the perihelion and in the
aphelion, which may be accounted for by varying the background rate and
the incoming neutrino flux. This explains the existence of islands.
For values of $\Delta m^2$ in between islands, the survival probability either
increases {\it and} decreases for varying Earth-Sun distance, or
monotonically {\it decreases}. The exact location of the islands and
their widths can only be understood by analyzing the fit procedure, in
particular the minimization of $\chi^2$ with respect to the background
rate and the incoming neutrino flux.
Note that there are no ``islands'' above $\Delta m^2\mathrel{\mathpalette\vereq>} 2.5\times
10^{-10}$~eV$^2$.
This is because when the oscillation length is small enough
(or $\Delta m^2$ large enough),
the survival probability {\it cannot} only increase for increasing
Earth-Sun distance, but necessarily reaches a maximum before the
aphelion, and then decreases, independent of what the survival
probability at the perihelion is. This situation is qualitatively
different from the low point.
In the case of the high point: initially the survival probability is
close to unity, decreases sharply as the Earth moves further from the Sun,
and then grows rapidly, reaching a maximum when the Earth is close to its
aphelion, because the oscillation length is smaller than
$\Delta L$. In this case, little
variations in the oscillation length, {\it i.e.} $\Delta m^2$,
produce big qualitative changes, including the position and number of
maxima and minima. There is still a small ambiguity ({\it i.e.}\/ two
``islands'') in determining $\Delta m^2$ for the high point. This happens
when the oscillation length is such that the minimum of the survival
probability happens in March/October and the survival probability is
large enough at the perihelion and the aphelion. The fact that the
absolute values of the number of recoil electrons detected are
different is taken care of by varying the signal and the
background.
In conclusion, if Nature chose neutrino oscillation parameters such
that $\sin^22\theta$ is large and $\Delta m^2\approx
\rm{few}\times10^{-10}$~eV$^2$, Borexino should be able to measure
these parameters independent of the SSM and any knowledge of the
number of background events, with good precision (especially in
$\Delta m^2$). If $\Delta m^2\approx
\rm{several}\times10^{-11}$~eV$^2$, the determination of oscillation
parameters is not as precise. Better precision can be achieved at
KamLAND, but the ambiguity of solutions in the ``low'' $\Delta m^2$
region still remains.
\setcounter{equation}{0}
\section{Exclusion of Vacuum Oscillations}
\label{sec:exclusion}
In this section, we address the issue of what the experiments can
conclude about vacuum
oscillations if no discrepancy from the normal seasonal variation
effect is detected. In this case, one may be able to measure the
incoming neutrino flux, as outlined in Sec.~\ref{sec:flux}. Two distinct
possibilities will be considered: (1) the measured flux is consistent
with the SSM prediction; (2) the measured flux is suppressed with
respect to the SSM prediction.
In the first case, one would be inclined to trust the SSM prediction
of the $^{7}$Be neutrino flux and use it in the analysis to exclude vacuum
oscillations. This will be discussed in Sec.~\ref{subsec:consistent}.
On the other hand, in the second case, it is not clear if the reduced
flux is due to MSW neutrino oscillations, an incorrect SSM prediction
of the neutrino flux, etc. This will be discussed in
Sec.~\ref{subsec:suppressed}.
\subsection{If the Flux is Consistent with the SSM Prediction}
\label{subsec:consistent}
We simulate ``data'' consistent with the SSM
and the expected number of background events. The relevant numbers are
quoted in Sec.~\ref{sec:flux}. The ``data'' are
binned into months (12 bins per year), and are illustrated
in Fig.~\ref{points}, assuming three years of Borexino running. We then fit
to the ``data'' annual distributions that include neutrino oscillations for
a given choice of $(\sin^22\theta,\Delta m^2)$, plus
a constant background. The background rate and the incoming neutrino
flux may be allowed to float in the fit, constrained to a positive number.
It is important to note that this is
the opposite of what was done in Sec.~\ref{sec:sensitivity}, where the
sensitivity of
Borexino and KamLAND to vacuum oscillations was studied. There, the
simulated ``data'' were consistent with vacuum oscillations, and one
tried to fit a non-oscillation prediction to the ``data'' by varying
the incoming flux and/or the background.
Here, the ``data'' are consistent with no oscillations, and one tries
to fit the ``data'' with a prediction which includes the effect of
neutrino oscillations
for fixed $(\sin^22\theta,\Delta m^2)$, by varying the incoming
flux and/or the background. If both the background and the incoming
flux are fixed, {\it i.e.}\/ not allowed to vary in the fit procedure,
the exclusion and the sensitivity regions are the same. On the other hand,
if both the background rate and the incoming flux are allowed to
float, the exclusion region is expected to be smaller than the
sensitivity region presented in Sec.~\ref{sec:sensitivity}, especially
in the region
$\Delta m^2\mathrel{\mathpalette\vereq<} 10^{-10}$~eV$^2$. This is due to the fact that a large
number of points in the parameter space yield an
annual variation of the $\nu_e$ flux which is much larger than
7\%, but agrees with the shape of the normal seasonal
variation. If in the fit procedure the signal is scaled down to reduce the
amplitude of the variation and the background scaled up to increase
the number of events, a good fit to the no oscillation case can be attained.
Fig.~\ref{exc_b} shows, for three years of Borexino and KamLAND
running, the region of the
$(\sin^2 2\theta,\Delta m^2)$ parameter space excluded at 95\% CL,
if one allows the solar neutrino flux and the
background rate to
float within the positive numbers (in white), and if one
assumes the solar neutrino flux calculated in the SSM within
theoretical errors (in light plus white).
\begin{figure}[tp]
\centerline{
\psfig{file=fig5_1.eps,width=1\textwidth}
}
\caption{Region of the two neutrino oscillation parameter
space excluded in the case of no neutrino oscillations if one
assumes no knowledge of the background and no knowledge of the SSM
(white) or knowledge of the SSM (light+white), after 3 years of
Borexino (right) and KamLAND (left) running.}
\label{exc_b}
\end{figure}
A few comments are in order. First, one notices that the KamLAND
exclusion region is larger than the one excluded by Borexino. This is,
of course, expected because of KamLAND's larger fiducial volume
and therefore higher statistics.
Second, when the solar neutrino flux is allowed to vary in the fit,
the excluded region of the parameter space shrinks, as expected and
discussed earlier.
Third, one can safely claim that, if no discrepancies are detected
in the seasonal variation spectrum, the ``large'' $\Delta m^2$ (several
$\times 10^{-10}$~eV$^2$) set of vacuum
solutions (see Figs.~\ref{fig:finalBorexino} and
\ref{fig:finalKamLAND}) will be excluded, even at
Borexino. Even when no knowledge of
the incoming neutrino flux is used, a reasonable portion
of the ``small'' $\Delta m^2$ (several $\times 10^{-11}$~eV$^2$) set of
solutions is also excluded. When one assumes knowledge of the incoming
neutrino flux, the entire allowed region is excluded.
If the background rate is larger than expected, the excluded region
diminishes accordingly. This is because when the constant background
is enhanced with respect to the oscillation signal it is easier
to achieve a reasonable $\chi^2$ for the fit even when the seasonal
variations due to vacuum oscillations are significantly different from the
no-oscillation case. In particular, when the background rate is large
enough that the seasonal distribution of the data is statistically
consistent with a flat one,
a reasonable $\chi^2$ for the fit can always be achieved simply by
scaling the signal to zero and scaling up the background
appropriately. Explicitly, after three years of Borexino (KamLAND)
running the exclusion region vanishes
if the background rate is
$\sim$8 (40) times larger than anticipated,
when both
the background rate and the incoming neutrino flux are allowed to
float in the fit or
$\sim$500 (3000) times larger than anticipated
when one assumes the neutrino flux predicted by the SSM.
\subsection{If There is an Overall Suppression of the Flux}
\label{subsec:suppressed}
If there is an overall, {\it i.e.}\/, time-independent suppression of the
flux (which is the case for the MSW solutions), the way to proceed towards
excluding part of the vacuum oscillation parameter space is less
clear. This is because such an experimental result neither agrees with
the SSM prediction nor does it represent any ``smoking gun'' signature
for neutrino oscillations, as is the case of anomalous seasonal
variations. One does not know if the SSM prediction of the flux
is simply wrong, or if there are neutrino oscillations consistent with
one of the MSW solutions or both. Anyway, it is clear that (in general) the
incoming neutrino flux should be considered unknown in the data analysis.
The most conservative option is to follow the same analysis done in the
previous subsection, and allow both the incoming neutrino
flux and the background rate to float in the fit. In this case, the
excluded region of the two-neutrino oscillation parameter space is reduced
significantly, and may completely disappear. This is because
when the number of signal events is reduced the annual distribution is
closer to flat and a good fit is obtained even when the would-be
annual variations are very different. This is very similar to what was
previously discussed at the end of the last subsection, where we
discussed what happens if the background rate turns out to be much larger than
anticipated. Explicitly, after three years of Borexino running and a
signal rate which is 21.3\% of the SSM prediction (as one would obtain
in the case of the small angle MSW solution), Borexino
is unable to exclude any portion of the vacuum
oscillation parameter space, while KamLAND can still exclude about one
half of the ``high'' and ``low'' $\Delta m^2$ preferred regions.
If the background rate can be estimated by other means with 10\%
uncertainty, Borexino and KamLAND will be able to exclude the entire``high''
$\Delta m^2$ region and a significant portion of the ``low''
$\Delta m^2$ region.
In order to go beyond the most conservative analysis discussed above,
one would have to look at the overall situation of the solar neutrino
puzzle at the time of the data analysis. It is likely that one will be able to do
much better.
For example, solar neutrino oscillations might have
already been established by the SNO experiment \cite{SNO}, and perhaps
it is reasonable to assume the incoming solar neutrino flux predicted
by the SSM.
Then it would be possible to exclude a region of the parameter space as large as
the one in Sec.~\ref{subsec:consistent} where one assumes the SSM flux.
Another possibility is that Super-Kamiokande or SNO rules out the
small angle MSW solution by studying the distortions of the
electron energy spectrum \cite{Super-K_spectrum,SNO}, and a large
suppression of the $^7$Be solar neutrino flux would indicate that there
is something wrong with the SSM. In this case, it is not clear how to
proceed. We do not go into further discussions on all
logical possibilities.
\setcounter{equation}{0}
\setcounter{footnote}{0}
\section{Conclusions}
\label{sec:conclusion}
We have studied possible uses of the seasonal variation of the $^7$Be
solar neutrino flux at Borexino and KamLAND. Our results can be
summarized as follows.
Once the experiments accumulate enough data to see seasonal
variations, the first step will be to determine if the observed pattern
is consistent with the normal $1/L^2$ flux suppression. If
a discrepancy is found, it will be a sign of vacuum
oscillations. In this case, the seasonal variation of the data can be
used to determine the oscillation parameters $\sin^22\theta$ and
$\Delta m^2$. On the other hand, if the data are consistent with the normal
pattern, the amplitude of the variation can be used to measure the
$^7$Be solar neutrino flux and to exclude a significant portion of
the vacuum oscillation parameter space.
If the observed seasonal variations are consistent with the
normal $1/L^2$ flux suppression, one can use the amplitude of the
variation to determine what fraction of the observed recoil electrons are
induced by the neutrinos coming from the Sun. This method is limited
by statistics, and the accuracy is worse when the $^7$Be solar
neutrino flux is suppressed, as in the case of the small angle MSW
solution. In fact, in Sec.~\ref{sec:flux} we found that in the case of a large
suppression only KamLAND should be able to perform such a measurement,
after 3 years of data taking. It is important to emphasize that
we assumed the oscillation of electron neutrinos into other active flavors. In the
case of oscillations into sterile neutrinos, the $^7$Be solar neutrino
flux might be almost absent, and in this case neither Borexino
nor KamLAND are able to perform a measurement of the flux using this technique.
An important advantage of this technique is that it does not
require a separate estimate of the background rate, which may be a
very difficult task. If the background rate
can be reliably measured by some other means, one can obtain
another measurement of the neutrino flux. In this case,
the two results can then be
compared for consistency, thus making the final result on the
$^{7}$Be neutrino flux much more
trustworthy.
We also studied in great detail the effect of vacuum neutrino
oscillations on seasonal variations.
Our analysis shows that the outlook for discovering vacuum oscillations at
both Borexino and KamLAND is very favorable.
A very important finding in Sec.~\ref{sec:sensitivity} is that the
experiments may detect
a deviation from
the normal pattern of seasonal variations even without relying on the
SSM prediction of the incoming neutrino flux or estimate of the
background rate.
The analysis would consist of trying to fit the observed data with the
normal $1/L^2$ pattern, treating the incoming neutrino flux and the
background rate as free parameters. With this technique, after three
years of running Borexino should detect anomalous seasonal variations for
almost all values of ($\sin^22\theta,\Delta m^2$) preferred by the
analysis of the neutrino flux data from Homestake, GALLEX, SAGE, and
Super-Kamiokande, as illustrated in
Fig.~\ref{fig:finalBorexino}. The sensitivity region should be
larger at KamLAND (Fig.~\ref{fig:finalKamLAND}).
Results obtained in this way would be very robust.
Both experiments are sensitive to an even larger portion of the parameter
space if the background rate can be reliably estimated by auxiliary measurements.
If anomalous seasonal variations are discovered, the data can be
used to measure the oscillation parameters ($\sin^22\theta$, $\Delta
m^2$). This issue was studied in Sec.~\ref{sec:measurement}. It was
found that for $\Delta
m^2 \mathrel{\mathpalette\vereq>} 10^{-10}$~eV$^2$ the experiments will be able to determine $\Delta
m^2$ with good precision. At the same time, for
$\Delta m^2 \mathrel{\mathpalette\vereq<} 10^{-10}$~eV$^2$ there would be many ``candidate
islands'' in the ($\sin^22\theta,\Delta m^2$) plane, and it will not
be easy to resolve the ambiguity.
On the other hand, the absence of anomalous seasonal variations of the
$^7$Be solar neutrino flux data can be used to exclude regions of the
vacuum oscillation parameter space. In Sec.~\ref{sec:exclusion} we
presented the
exclusion plots for both Borexino and KamLAND, after three years of
running. An important lesson
from that section is that in order to exclude a large portion of the
preferred region, the
experiments will need to either measure the background rate or rely on
the SSM prediction for the neutrino flux. In the absence of both, the results are
rather weak. This is to be contrasted with the situation in
Sec.~\ref{sec:sensitivity}.
It is important to keep in mind that the simulated ``data'' is most of
the time based on the
SSM prediction for the $^7$Be solar neutrino flux and the anticipated
number of background events at Borexino and KamLAND.
Our {\it numerical} results, therefore, even in the cases when we do not
use the knowledge of the incoming neutrino flux or the background rate
{\it at the analysis stage}, are not to be regarded as SSM and
background rate independent.
We would like to draw attention to our comments at the end of
Secs.~\ref{sec:flux} and \ref{sec:sensitivity} on how our results
might change if these
inputs are changed. We also assume only statistical errors in the data
analysis, neglecting systematic uncertainties due to the lack of
knowledge in the seasonal variation of the background rate. The inclusion of such
effects is beyond the scope of this paper.
Overall our results indicate that the future Borexino results can
lead to significant progress towards solving the solar neutrino
puzzle. Furthermore, if KamLAND is also able to study solar neutrinos,
one would have access to
a larger data set, and more powerful results can be obtained.
\section*{Acknowledgments}
We would like to express special thanks to Eligio Lisi and Lincoln Wolfenstein
who pointed out the irrelevance of the core size effect.
HM also thanks John Bahcall,
Stuart Freedman, and Kevin Lesko for useful conversations. This work
was supported in part by the U.S. Department of Energy under Contracts
DE-AC03-76SF00098, in part by the National Science Foundation under
grant PHY-95-14797. HM was also supported by Alfred P. Sloan
Foundation, and AdG by CNPq (Brazil).
|
\section{Introduction}
\label{introduction}
Neutrino physics is one of the most active fields in particle
physics nowadays. Apart from the impressive results of the
underground experiments concerning atmospheric and solar
neutrino measurements, also reactor and accelerator physics have
a large share in the evolution of this field and the
understanding of the neutrino mass spectrum and mixing matrix
(for recent reviews see, e.g., Refs. \cite{conrad,BGG98}). Since
the phenomenon of neutrino oscillations is now close to being an
established physical reality and constitutes the most tangible
window for physics beyond the Standard Model, it is very important
to know if there are any limitations to the validity of the usual
formula for neutrino survival and
transition probabilities \cite{pon} with which the experimental results
are evaluated. Such questions have been discussed extensively in
the literature in the context of the wave packet and
field-theoretical approaches (see, e.g., Ref. \cite{zralek} for a list of
references).
In this paper we use the field-theoretical approach in the spirit of
Ref. \cite{rich},
which has proved to be a general and unambiguous method to analyse
neutrino oscillations, and concentrate on the effect of the finite
lifetime of the neutrino source particle.
We study the problem under the following three assumptions:
\begin{enumerate}
\renewcommand{\labelenumi}{\roman{enumi}.}
\item the neutrino source is at rest,
\item the source particle is not in a bound state and, therefore,
is described by a non-stationary wave function,
\item the wave function of the detector particle is stationary.
\end{enumerate}
The first assumption is of technical nature and allows us to use
the methods and results of Refs. \cite{GS96,GSM99} and can
probably be overcome \cite{campagne}, however, the second one
is essential to the discussion in this paper.
The third assumption simply means that the particle with which
the neutrino reacts in the detector is in a bound
state and the wave function of the detector particle does not spread with
time \cite{GSM99}. Note that we also neglect for both the production
and the detection process any possible interaction with the background.
The finite lifetime $\tau_S = 1/\Gamma$ of the neutrino
source particle has an impact on the neutrino oscillation
probability in the following two ways:
\begin{enumerate}
\renewcommand{\labelenumi}{(\alph{enumi})}
\item a suppression of the probability
amplitude as a function of $\Gamma$ \cite{GSM99}, an effect
independent of $L$, the distance between neutrino source and detection,
\item a damping of neutrino oscillations through the coherence
length $L^\mathrm{coh}_\Gamma$ caused by the finite lifetime and
given by the factor $\exp (-L/L^\mathrm{coh}_\Gamma)$ \cite{GSM99} in
the oscillation probability.
\end{enumerate}
The main topic of this paper is the derivation of a correction
to the usual neutrino oscillation probability according to point
(a) and a numerical study of this correction. It has been
shown in Ref. \cite{GSM99} that the effect (a) of $\Gamma$
is only present provided assumption ii) is valid.
This seems to be the case in the important experiments of the
LSND \cite{LSND} and KARMEN \cite{KARMEN} Collaborations,
which study $\bar\nu_\mu\to\bar\nu_e$ oscillations with
$\mu^+$ decay at rest as $\bar\nu_\mu$ source.\footnote{In the LSND
experiment also $\nu_\mu\to\nu_e$ oscillations are investigated where
the muon neutrinos originate from $\pi^+$ decay in flight. According
to assumption i) we do not discuss this case here.} In the
following we will use the source and detection reactions of LSND
and KARMEN as a model for our field-theoretical treatment of
neutrino oscillations.
The paper is organized as follows.
In Section II we derive the transition probability
for $\bar\nu_\mu\rightarrow\bar\nu_e$ oscillations and show how the
finite source lifetime changes the standard formula of
neutrino oscillations and gives rise to a suppression factor for the
interference terms in the oscillation formula. We will see that the
corresponding condition for the existence of neutrino oscillations,
i.e., the condition that the suppression does not take place, depends
only on parameters supplied, at least in principle, by the experimental
set-up and the neutrino mass squared difference $\Delta m^2$. This is
natural within the field-theoretical treatment which enables the
study of the dependence on those quantities which are really observed
or manipulated in oscillation experiments.
In Section III we apply the
results of Section II to the case of a Gaussian $\mu^+$ wave
function and make a numerical study of the 2-flavour transition
probability to investigate quantitatively the influence of the
correction factor. In Section IV we present a discussion of our
findings and draw the conclusions.
\section{The cross section}
As mentioned in the introduction, for definiteness and also for
comparison with the notation in Ref. \cite{GSM99} we discuss the process
\begin{equation}\label{process}
\mu^+ \to e^+ + \nu_e + \bar\nu_\mu
\stackrel{\nu\: \mathrm{osc.}}{\leadsto}
\bar\nu_e + p \to n + e^+ \,,
\end{equation}
which is investigated in the LSND and KARMEN experiments.
As shown in Ref. \cite{GSM99}, with the help of the Weisskopf--Wigner
approximation one can write the amplitude of
the process (\ref{process}) in the limit $t \rightarrow \infty$ as
\begin{eqnarray}
{\cal A} & = & (-i)^2
\langle \nu_e (p'_\nu), e^+_S (p'_{eS}); e^+_D (p'_{eD}), n (p'_n) | \nonumber\\
&& \times T \left[ \int_0^\infty \! dt_1 \int d^3x_1
\int_0^\infty \! dt_2 \int d^3x_2\, {\cal H}^+_{S,\mathrm{int}}(x_1)
e^{-\frac{1}{2}\Gamma t_1} {\cal H}^+_{D,\mathrm{int}}(x_2)
\right] |\mu^+ ; p \rangle \,,
\label{a}
\end{eqnarray}
where $T$ is the time-ordering symbol and $\Gamma$ is the total decay width
of the muon. ${\cal H}^+_{S,\mathrm{int}}$ and
${\cal H}^+_{D,\mathrm{int}}$
are the relevant Hamiltonian densities (in the interaction picture)
for the production and
the detection of the neutrinos, describing muon decay and proton to neutron
transition (\ref{process}), respectively.
The indices $S$ and $D$ denote source and detection, respectively.
The muon $\mu^+$ and the proton $p$ are localized at the
coordinates $\vec x_S$ and $\vec x_D$, respectively.
The proton state is stationary whereas the decaying muon
at rest is described by a free wave packet with an average momentum equal
to zero according to the assumptions i)--iii). Hence the
spinors of the initial particles are written as
\begin{eqnarray}
\Psi_p (x) = \psi_p (\vec x - \vec x_D) \, e^{-i E_p t}
& \quad \mbox{with} \quad &
\psi_p (\vec x-\vec x_D) = \frac{1}{(2\pi)^{3/2}} \int d^3p\,
\widetilde{\psi}_p(\vec p)\, e^{i\vec p \cdot (\vec x-\vec x_D)}
\nonumber\\
& \quad \mbox{and} \quad &
\widetilde{\psi}_p(\vec p) = \widetilde{\psi}^\prime_p(\vec p)
u_p(\vec p)
\label{proton}
\end{eqnarray}
for the proton and
\begin{eqnarray}
\psi_\mu (x) = \int \frac{d^3 p}{(2\pi)^{3/2}}\, \widetilde{\psi}_\mu
(\vec p)\, e^{- i (\vec p \cdot \vec x - E_\mu(\vec p)t)} \times
e^{i \vec p \cdot \vec x_S}
& \quad \mbox{with} \quad &
\widetilde{\psi}_\mu(\vec p) = \widetilde{\psi}^\prime_\mu(\vec p)
v_\mu(\vec p)
\end{eqnarray}
for the muon with $E_\mu(\vec{p}) = \sqrt{m^2_\mu+\vec{p}\,^2}$.
In the 4-spinors $u_p(\vec p)$ and $v_\mu(\vec p)$
we have left out the polarizations of the proton and muon,
respectively, because they are irrelevant in the further discussion.
The function $\psi_p (\vec y)$ is peaked at $\vec y = \vec 0$
and the wave packet $\widetilde{\psi}_\mu (\vec p)$
in momentum space is peaked
around the average momentum $\langle \vec p\rangle = \vec 0$.
The final particles are described by plane waves.
After carrying out all integrations, the leading term of the
amplitude in the asymptotic limit $L \to \infty$
can be written as \cite{GSM99}
\begin{equation}\label{ampinfty}
{\cal A}^\infty = \sum_j U_{\mu j}U^*_{ej} e^{i q_j L}
{\cal A}^S_j {\cal A}^D_j \,,
\end{equation}
where ${\cal A}^S_j$ and ${\cal A}^D_j$ denote the amplitudes for
production and detection (\ref{process}) of a neutrino with mass
$m_j$, respectively. The parts of these amplitudes which are important for the
further discussion are given by
\begin{equation}\label{as}
{\cal A}^S_j =
\frac {1}{i (E_{Sj} -E_D) + \frac{1}{2}\Gamma}\,
\overline{\widetilde{\psi}}_\mu (\vec p_1 + q_j \vec l\,) \,\cdots
\end{equation}
and
\begin{equation}\label{ad}
{\cal A}^D_j = \cdots\,
\widetilde{\psi}_p(-q_j \vec l + \vec p_2) \,.
\end{equation}
For the full expressions see Ref. \cite{GSM99}.
The kinematical quantities occurring in Eqs.~(\ref{as}) and (\ref{ad})
are defined by
\begin{eqnarray}\label{def1}
q_j & = & \sqrt{E_D^2 - m_j^2} \,, \nonumber\\
E_D & = & E'_n + E'_{eD} - E_p \,, \nonumber\\
E_{Sj}& = & E_\mu (q_j \vec l + \vec p_1) - E'_\nu - E'_{eS}
\end{eqnarray}
and
\begin{equation}\label{def2}
\vec p_1 = \vec p\,'\!\!_\nu + \vec p\,'\!\!_{eS} \,,\quad
\vec p_2 = \vec p\,'\!\!_n + \vec p\,'\!\!_{eD}
\quad \mbox{and} \quad
L = | \vec x_D - \vec x_S | \,.
\end{equation}
Note that the first two formulas in Eq.~(\ref{def1}) follow from
assumption iii) in the introduction \cite{GS96,GSM99}.
Eq.~(\ref{as}) shows the dependence of ${\cal A}^\infty$ on $\Gamma$
(see point (a) in the introduction).
The structure (\ref{ampinfty}) arises from the fact that in
the limit $L\rightarrow\infty$ the neutrinos with mass $m_j$, described
by an inner line of the Feynman diagram derived from ${\cal A}$
(\ref{a}), are on mass shell. Consequently, the
amplitude for neutrino production and its subsequent detection factorizes
for each $j$ into a product of production amplitude and detection amplitude
\cite{GS96}.
Looking at Eqs.~(\ref{ampinfty}), (\ref{as}) and
(\ref{ad}) it is evident that oscillations
involving $m^2_j-m^2_k$ can only take place if the
conditions \cite{rich,GS96}
\begin{equation}\label{ACC}
|q_j-q_k| \lesssim \sigma_S \quad \mbox{and} \quad
|q_j-q_k| \lesssim \sigma_D
\end{equation}
and \cite{GSM99}
\begin{equation}\label{SFC}
|E_{Sj}-E_{Sk}| \lesssim \frac{1}{2} \Gamma
\end{equation}
hold, where $\sigma_S$ and $\sigma_D$ are the widths of
$\widetilde{\psi}_\mu$ and $\widetilde{\psi}_p$, respectively.
We call conditions (\ref{ACC}) amplitude coherence conditions (ACC)
and equation (\ref{SFC}) the source wave packet -- finite lifetime
condition (SFC) \cite{GSM99}.
If they are not fulfilled then some terms of the amplitude and
consequently the corresponding
interference terms in the cross section (oscillation probability)
are suppressed.
The cross section $\sigma$ is obtained by taking the absolute
square of the amplitude (\ref{ampinfty}),
integrating over the final state momenta $\vec p\,'\!\!_\nu$,
$\vec p\,'\!\!_{eS}$,
$\vec p\,'\!\!_n$, $\vec p\,'\!\!_{eD}$ and averaging over the spins
of the initial muon and proton. Hence
\begin{equation}\label{sigma}
\sigma = \int dP\; |{\cal A}^\infty|^2 \,,
\end{equation}
where $dP$ denotes the integration over the momenta of the particles
at the external legs and is given by
\begin{equation}
dP = \frac{d^3p'_\nu}{2E'_\nu}\; \frac{d^3p'_{eS}}{2E'_{eS}}\;
\frac{d^3p'_n}{2E'_n}\; \frac{d^3p'_{eD}}{2E'_{eD}}
\end{equation}
and the integration is done over some volume of the phase space.
In general, the integrations in (\ref{sigma})
cannot be performed without knowledge of the source and detector wave
functions. However, as noticed in Ref. \cite{GSM99}, for
\begin{equation}\label{IC}
\Gamma \ll \sigma_{S,D}
\end{equation}
the factors
\begin{equation}\label{factor1}
\left\{(i(E_{Sj}-E_D)+\Gamma/2) \times
(-i(E_{Sk}-E_D)+\Gamma/2)\right\}^{-1}
\end{equation}
in the cross section are strongly peaked with respect to $E_D$
and we interpret the conditions (\ref{IC}) that the rest of the
cross section is flat with respect to $E_D$ if varied over intervals several
orders of magnitude larger than $\Gamma$. This is a reasonable
assumption because in the LSND and KARMEN experiments the
stopped muons have momenta of the order 0.01 MeV \cite{louis,drexlin}, and
thus $\sigma_S$ will be in the same range. Assuming
atomic dimensions of the spread of the detector particle wave
function, we find $\sigma_D \sim 10^{-3}$ MeV. In any case, even
if our guesses for $\sigma_S$ and $\sigma_D$ are wrong by
several orders of magnitude, these widths can never be so small
like the decay constant of the muon $\Gamma \simeq 3 \times
10^{-16}$ MeV. Therefore, we adopt the procedure that
integrating over momenta of the final state of the detector
leads to an integration in the variable $E_D$ and in view of the
expressions (\ref{factor1}) it suffices to pick out an
integration interval of length $\Delta E_D$ which fulfills
$\Gamma \ll \Delta E_D \ll \sigma_{S,D}$. Considering neutrino
oscillations with respect to $m^2_j - m^2_k \equiv \Delta m^2_{jk}$,
then this integration interval should contain
$E_{Sj}$ and $E_{Sk}$. Since the condition (\ref{SFC}) is
necessary for these oscillations to happen at all, we define a
mean value
\begin{equation}\label{dED}
\bar{E}_D \equiv \langle E_{Sj} \rangle
\end{equation}
for the neutrino mass eigenfields participating in the
oscillations and $\Delta E_D$ only has to be a few orders of
magnitude larger than $\Gamma$. In addition, the expressions
(\ref{factor1}) suggest to close the integration over the
interval of length $\Delta E_D$ along the real axis by a half
circle in the complex plane and to apply Cauchy's theorem \cite{GSM99}.
In this way we have shown that the cross section contains the
damping factor \cite{GSM99}
\begin{equation}\label{damping}
\exp \left( - \frac{\Delta m^2_{jk}\Gamma}{4\bar{E}^2_D}\,L \right)
\end{equation}
leading to a suppression
of the corresponding interference term in the oscillation
probability when $L$ becomes greater than the coherence length
$L^\mathrm{coh}_\Gamma = 4 \bar E_D^2/\Delta m^2_{jk} \Gamma$.
As already discussed in Ref.
\cite{GSM99}, this coherence length is of the order of 100 light years
for typical input relevant for the LSND and KARMEN experiments
and its effect is certainly negligible for all experiments
with terrestrial neutrinos. In the following computations we
will neglect the coherence length arising from the finite lifetime of the
neutrino source.
Then, the only relevant integrand with respect to $E_D$ is
given by the factors (\ref{factor1}) and we get
\begin{equation}\label{int1}
\int_{-\infty}^\infty dE_D\,\frac{1}{i(E_{Sj}-E_D)+\frac{1}{2}\Gamma}\,
\frac{1}{-i(E_{Sk}-E_D)+\frac{1}{2}\Gamma} = \frac{2\pi}{\Gamma}\,\frac{1}
{1+i\frac{1}{\Gamma}(E_{Sj}-E_{Sk})} \,.
\end{equation}
On the right-hand side of this equation we can use assumption
i) (the muon at rest) to compute to a very good approxmimation
the energy difference
\begin{equation}
E_{Sj}-E_{Sk}\simeq -\frac{\Delta m^2_{jk}}{2m_\mu \bar E_D}\,\vec{l}
\!\cdot\!(\vec{p_1}
+\bar E_D \vec{l}\,) \,.
\end{equation}
As the next step, we want to perform the integration over
$d^3p'_\nu$. Note that the momentum $\vec p\,'\!\!_\nu$ of
the $\nu_e$ from the source reaction (see Eq.~(\ref{process}))
cannot be measured.
Again we make a simplifying approximation based on assumption i),
namely that the muon wave function can be represented by
\begin{equation}
\widetilde{\psi}_\mu(\vec{p}_1+q_j\vec{l}\,) \simeq
\widetilde{\psi}'_\mu(\vec{p}_1+\bar E_D \vec{l}\,)\,
v_\mu(\vec{0}\,) \,.
\end{equation}
Then the integration over
$d^3p'_\nu$ concerns only the functions $(E_{Sj}-E_{Sk})$
in (\ref{int1}) and
$\widetilde{\psi}^\prime_\mu$. Therefore, we define the quantities
\begin{equation}\label{gfact}
g_{jk}= \int d^3u\, |\widetilde{\psi}'_\mu(\vec{u})|^2
\frac{1}{1-i\rho_{jk}\,\vec{l}
\!\cdot\! \vec{u}/\sigma_S}\, ,
\end{equation}
where we have defined $\vec{u}\equiv \vec{p}_1+E_\nu\vec{l}$ and
\begin{equation}\label{rho}
\rho_{jk}\equiv\frac{\Delta m^2_{jk}\sigma_S}{2 m_\mu E_\nu\Gamma} \,.
\end{equation}
Note that in view of Eqs.~(\ref{def1}) and (\ref{dED}) we identify
the neutrino energy which can in principle be measured by the
detector as
\begin{equation}\label{eneu}
E_\nu\equiv\bar E_D \,.
\end{equation}
The quantities (\ref{gfact}) have the properties that
$g_{jk}=g_{kj}^*$ and $g_{jj}=1$. Since we consider
ultrarelativistic neutrinos, we can take the limit
$m_j \rightarrow 0$ $\forall \,j$ in all terms of the cross
section (\ref{sigma}) except the oscillation phases and
$g_{jk}$. Then the probability for $\bar\nu_\mu\to\bar\nu_e$
oscillations is given by
\begin{eqnarray}
P_{\bar\nu_\mu\to\bar\nu_e} &=&
\sum_j |U_{e j}|^2 |U_{\mu j}|^2 +\nonumber\\
&&2\, \mbox{Re} \left\{ \sum_{j>k}
U^*_{e j} U_{\mu j} U_{e k} U^*_{\mu k}\,\,
g_{jk}\,\exp \left( -i \frac{\Delta m^2_{jk} L}{2E_\nu} \right)\,
\right\} \,.
\label{P}
\end{eqnarray}
In contrast to the formula derived within the standard
treatment of neutrino oscillations, the quantities $g_{jk}$ appear
as correction factors in
Eq.~(\ref{P}). These non-standard factors $g_{jk}$
represent the quantitative effect of the SFC condition (\ref{SFC})
discussed in Ref. \cite{GSM99}. They arise from assumption ii)
(see introduction) that the muon is described by a free wave function,
which does not have a sharp energy, and would disappear for a muon
in a stationary (bound) state (see Ref. \cite{GS96}).
\section{The case of a Gaussian muon wave packet}
In order to get a feeling for the effect of the $g$-factors (\ref{gfact})
on the oscillation probabilities we confine
ourselves now to the case of a Gaussian muon wavefunction
\begin{equation}\label{gauss}
\widetilde{\psi}^\prime_\mu(\vec p) = (\sqrt{ \pi} \sigma_S)^{-3/2}
\exp{\left(-\frac{\vec p\,^2}{2\sigma_S^2}\right)} \,.
\end{equation}
Then we get
\begin{equation}\label{ggfact}
g_{jk}=\frac{1}{\sqrt{\pi}\sigma_S}\int_{-\infty}^\infty du\,
e^{-u^2/\sigma^2_S} \frac{1}{1-i\rho_{jk} u/\sigma_S} \,,
\end{equation}
where $u\equiv \vec{l}\!\cdot\!\vec{u}$.
Since the exponential factor in Eq.~(\ref{ggfact}) is an even
function with respect to the variable $u$ the $g$-factors become
real and defining $y = u/\sigma_S$ one obtains $g_{jk} = g(\rho_{jk})$
with
\begin{equation}\label{g}
g(\rho) \equiv
\frac{1}{\sqrt{\pi}}\int_{-\infty}^\infty dy \frac{e^{-y^2}}{1+
\rho^2 y^2}
=\frac{\sqrt{\pi}}{\rho} \exp \left( \frac{1}{\rho^2} \right)
\Phi_c \left( \frac{1}{\rho} \right) \,,
\end{equation}
where $\Phi_c$ is the complementary error function (see, e.g., Ref.
\cite{Gradst}). Note that
$g(\rho) \simeq 1-\rho^2/2$ for $\rho \ll 1$
and $g(\rho) \to \sqrt{\pi}/\rho$ for $\rho \to \infty$.
Under the assumption of oscillations between two neutrino flavours
the transition probability becomes
\begin{equation}\label{gProb}
P_{\bar\nu_\mu\rightarrow\bar\nu_e}=\frac{1}{2}\sin^2 2\theta
\left( 1-g_{12} \cos\frac{\Delta m^2 L}{2 E_\nu} \right) \,,
\end{equation}
where $\theta$ is the 2-flavour mixing angle. In comparison with
the standard
formula the factor $g_{12}$ in (\ref{gProb}) lowers for a given transition
probability the upper bounds on $\Delta m^2$ and $\sin^22\theta$.
This can be seen explicitly by considering
the low $\Delta m^2$-region of the parameter
space where the cosine in expression (\ref{gProb}) can be
expanded in $\Delta
m^2$. If now $\rho\lesssim 1$ holds for an average $\Delta m^2$, in the
low $\Delta m^2$-region we get $\rho\ll 1$ and can use the corresponding
expansion of $g(\rho)$ given above.
Then, in this region the isoprobability contour is determined
by the relation
\begin{equation}
(\Delta m^2 )^{2}\,\sin^2 2\theta \simeq
{16E_\nu^2\, P_{\bar\nu_\mu\rightarrow\bar\nu_e} \over
L^2 + (\frac{\sigma_S\tau_S}{m_\mu})^2}\, .
\label{isop}
\end{equation}
It is clear from Eq.~(\ref{isop}) that a large source lifetime and a large
source momentum spread lower the upper bounds on $\Delta m^2$
and $\sin^2 2\theta $ compared to the standard oscillation formula. For the
source lifetime effect to be significant in a given experiment,
the numerical value of $\sigma_S\tau_S/m_\mu$ must be
comparable to the source detector distance $L$.
In the following we investigate quantitatively the effect of $g_{12}$
on the transition probability (\ref{gProb}). In Fig.~1 we display
the function $g(\rho)$. One can see that at $\rho = 0.5$
the function $g$ is around 0.9 which means that $g_{12}$
starts to deviate appreciably from 1 when $\rho$ becomes greater than
about 0.5. In the case of the LSND and KARMEN experiments, using
$\Gamma \simeq 3\times 10^{-16}$ MeV for the decay width of the muon
and the typical numbers $\sigma_S \simeq 0.01$ MeV \cite{louis,drexlin},
$\Delta m^2 \simeq 1$ eV$^2$ and $E_\nu \simeq 30$ MeV, we obtain
$\rho_{12}\simeq 0.5\times 10^{-2}$.
From this estimate we conclude that with the above input numbers
the correction to the transition probability is negligible for
the LSND and KARMEN experiments because
$1-g_{12} \sim 10^{-5}$ (see remark after definition of $g$ (\ref{g})).
Note, however, that $\rho_{12}\simeq 0.5\times 10^{-2}$ is only
two orders of magnitude away from having a 10\% effect.
This motivates us to have a look at the influence of $\rho_{12}$ ($g_{12}$)
on exclusion curves obtained, for instance, in the KARMEN
experiment.
Therefore, we use $L = 17.7$ m, average in Eq.~(\ref{gProb}) over the energy
spectrum of $\bar{\nu}_\mu$ and as a Gedankenexperiment we vary the momentum
spread $\sigma_S$.
In Fig.~2 we show the corresponding exclusion curves in the $\Delta m^2$
-- $\sin^2 2\theta$ plane for the cases of
$\sigma_S=0$, 1 and 10 MeV corresponding to $g_{12} \simeq
1$, 0.9 and 0.3, respectively.
We can see that for $\sigma_S=10$ MeV the exclusion curve has changed
noticeably.
\section{Discussion and conclusions}
In this paper we have studied the effects of the finite lifetime of
the neutrino source particle. In particular, we have eleborated
the condition for the existence of
neutrino oscillations arising in a situation where the neutrino
source particle is not described by a bound state and thus
represented by a non-stationary wave function.
Furthermore, in our analysis we have
made the assumption that the neutrino source is at rest.
In order to perform an analysis closely related to experiment we
used the field-theoretical approach which allows to work directly with
the states of the particles responsible for neutrino production and
detection
and where the oscillating neutrinos are represented by an inner line
in the complete Feynman diagram containing the source and detection
processes.
Thus the neutrino oscillation amplitude is determined simply through
the neutrino production and detection interaction Hamiltonians.
For simplicity, in the following we will concentrate on 2-flavour
oscillations.
The finite lifetime of the neutrino source leads to a finite coherence
length \cite{KNW96}. In our case we have explicitly indicated the
damping factor (\ref{damping}) from where the coherence length
$L^\mathrm{coh}_\Gamma$ associated with $\Gamma$ can be read
off. Using numbers taken from the LSND and KARMEN experiments for the
quantities appearing in $L^\mathrm{coh}_\Gamma$ we get a length of the
order of 100 light years \cite{GSM99} which is absolutely negligible
compared the the coherence length stemming from the fact that the
neutrino energy $E_D$ (see Eqs.~(\ref{def1}) and (\ref{dED}))
determined by the energies
of the final states in the detection process is not accurately
known. Therefore, the effect (b) of $\Gamma$ (see introduction) is
irrelevant.
Let us now compare the amplitude coherence condition (ACC) with the source
wave packet -- finite lifetime condition (SFC). The ACC can be
reformulated as \cite{kayser,rich,GS96}
\begin{equation}\label{ACC1}
\sigma_{xS,\, xD} \lesssim \frac{1}{4\pi} L^\mathrm{osc} \, ,
\end{equation}
where $\sigma_{xS,\, xD}$ are the widths of the source and detector
wave functions, respectively, in coordinate space and
$L^\mathrm{osc}=
4\pi E_\nu/\Delta m^2=2.48\; \mbox{m}\; (E_\nu/1\;\mbox{MeV})
(1\;\mbox{eV}^2/\Delta m^2)$ is the oscillation length.
For the LSND and KARMEN
experiments with $E_\nu\sim30$ MeV and $\Delta m^2\sim 1\;\mbox{eV}^2$,
as found by the LSND collaboration,
one gets $L^\mathrm{osc}\sim 75$ m. This is to be compared
with the widths $\sigma_{xS,\, xD}$ which are certainly in the
range of atomic distances or smaller. Therefore the conditions
(\ref{ACC1}) are very well fulfilled.
We want to stress once more that the SFC and the correction factor $g_{12}$
(\ref{gfact}) corresponding to effect (a) of $\Gamma$ (see
introduction) derive from two ingredients:
from the finite lifetime of the muon taken into account through the
Weisskopf--Wigner approximation and from
assumption ii) in the introduction that the wave function of the muon
is non-stationary. The first ingredient leads to
the factor $1/(i(E_{Sj}-E_D)+\frac{1}{2}\Gamma)$ in the amplitude instead
of the familiar delta function. The second one leads to the momentum
dependence of $E_{Sj}$ as given in Eq.~(\ref{def1}). If the neutrino
source particle were in a bound state and, therefore, described by a
stationary wave function, $E_S$ would be fixed and independent of $j$
and momenta, and there would be no condition (\ref{SFC}). Note that in
our approximation the Hamiltonian which determines
the wave function of the source particle does not
include weak interactions. However, the decay of the source particle
via the weak interactions -- whether its wave function is stationary or
non-stationary -- is explicitly given by the factor
$\exp (-\Gamma t_1/2)$ in the amplitude (\ref{a}) according to the
Weisskopf--Wigner approximation.
As shown in Section III the correction factor $g_{12}$ depends on
the quantity
\begin{equation}\label{rho1}
\rho = \frac{\Delta m^2 \sigma_S}{2 m_\mu E_\nu\Gamma}\; ,
\end{equation}
and is only sizeably different from one if $\rho \gtrsim 1$. Thus
a convenient formulation of SFC is $\rho < 1$.
Note that by defining a spread in velocity of the source
wave packet by $\Delta v_S = \sigma_S/m_\mu$ the last formulation of
SFC can be rewritten in analogy to Eq.~(\ref{ACC1}) as \cite{GSM99}
\begin{equation}\label{SFC1}
\Delta v_S \tau_S \lesssim \frac{1}{4\pi} L^\mathrm{osc} \,.
\end{equation}
As we have seen in
Section III with $\sigma_S \sim 0.01$ MeV one gets $\rho \sim 0.005$
and SFC is well fulfilled for the LSND and KARMEN experiments,
however, the margin for violation of SFC is only two orders of
magnitude, in contrast to ACC where the margin is at least ten orders
of magnitude.
Taking this observation as a starting point,
we want to see if there is
some chance to increase $\rho$ to 1 or more in order to have an
observable suppression effect.
Looking at Fig.~2 one recognizes that in the case of
$\sigma_S=10$ MeV the lower line of the curve has undergone a
considerable shift downwards compared to the case $\rho = 0$.
Note that this shift happens in
the sensitive region of the LSND and KARMEN experiments. One might
therefore ask the question if it is possible to increase $\sigma_S$ in
order to achieve $\rho \gtrsim 1$. If
the source wave function is such that $\sigma_S\, \sigma_{xS} \sim 1$,
which is correct for Gaussian wave functions,
one would have to localize the neutrino
source particle very well in coordinate space, namely to $\sigma_{xS}
\lesssim 10^{-12}$ m
for $\sigma_S \gtrsim 0.1$ MeV.
Taking a look at Eq.~(\ref{rho1}), one
observes that $\rho$ can also be increased by decreasing the neutrino
energy. However, in this case one lowers the ratio of true over
background events.
Finally, a small product of mass times decay width of the
source
particle also enhances $\rho$ but, unfortunately, the muon
has already
the smallest such product among the particles which can
copiously be
produced.\footnote{The corresponding
product for the neutron is smaller, however, to observe this effect in
neutrinos from neutron decay one would have to isolate the neutrons from the
external interactions -- something that seems to be hard to achieve in
experiments.}
The quantum-mechanical ACC and SFC in the forms (\ref{ACC1}) and
(\ref{SFC1}), respectively, have analogous ``classical
macroscopic'' conditions stemming from the inescapable averaging over some
regions of the target and the detector, respectively, when an
experiment is performed.
Replacing in Eq.~(\ref{ACC1}) the widths $\sigma_{xS,\, xD}$
by the typical sizes $R_{S,D}$, respectively, of the regions in the target
where the muon is stopped and in the detector where the neutrino detection
process is localized we arrive at conditions which are obtained by the
incoherent averaging of the oscillation probabilities over the variations
of $L$ due to $R_{S,D}$ and the requirement that
neutrino oscillations should not be washed out by this averaging process.
Since $R_{S,D}$ are macroscopic quantities,
if these classical conditions are fulfilled, then clearly also ACC
holds because $\sigma_{xS,\, xD} \ll R_{S,D}$. The classical analogue
to SFC (\ref{SFC1}) says that during its lifetime the
neutrino source particle should move a distance much
less than the oscillation length in order not to wash out neutrino
oscillations. This classical condition is not obviously linked to SFC.
|
\section{Introduction}
In the semiclassical treatment of quantum field theories,
instantons \cite{BPST,tH,dilute}
play the important role of providing the tunneling between
topologically inequivalent
vacuua in an essentially nonperturbative framework. Instantons
are classical solutions
to the Euler-Lagrange equations with Euclidean time, whence comes
the tunneling
interpretation. Another, closely related mechanism
for vacuum to vacuum transitions but now at non--zero temperature, is
provided by the
(Euclidean) static solutions of the field equations called
sphalerons \cite{Manton}. In
contrast to instantons, sphalerons are unstable solutions and
provide a classical
rather than quantum (tunneling) transition over the energy barrier
separating the two
vaccua. In this context it is very useful to consider another class of
classical
solutions, which are periodic in (Euclidean) time, with the period being
identified with the inverse of the temperature. These are the periodic
instantons
as defined in Ref.~\cite{KRT}. Thus at zero temperature the period of
the periodic
instanton becomes infinite, rendering it non-periodic, which can be
identified as the
instanton itself.
Now a given field theoretic model supporting sphalerons
\cite{Manton,Klinkhamer} and
hence also
periodic instantons \cite{KRT}, may or may not support (zero--temperature or
infinite period) instanton. In
the case where a finite--size instanton exists, it can
be arrived at as the period of
the periodic instanton tends to infinity. In the case however
where no finite size
instanton exists, the situation is more complicated and the
so--called constrained
instantons \cite{A} must be employed. The consequences of a
theory being of one type
or the other are thought to be potentially important.
There has been some exploratory work done in this direction,
in the context of the $1+1$
dimensional scale--breaking $O(3)$ sigma model \cite{MW} which
does not support finite
size instantons, and its skyrmed version \cite{PMTZ} which
does so. The study of the
periodic instantons in these two models from this viewpoint
was carried out in refs.~\cite{HMT,KT} respectively.
In view of the above, it is interesting to construct and
study models that can support
both {\it instantons} and {\it sphalerons}, as a first
step before studying the {\it periodic
instantons} interpolating them. Our aim in this paper
is to do just this, for a class of
$d$-dimensional $SO(d)$ Higgs models in which the
Higgs field is a $d$-component vector of
the $SO(d)$, for the two cases of $d=2$ and $d=3$.
What distinguishes such models is that
their instantons result in curvature field strengths
that exhibit {\it inverse square}
behaviour. This property contrasts with the pure-gauge
behaviour of the arbitrary--scale
Yang-Mills (YM) instantons \cite{BPST} and can result
in far reaching (physical) consequences.
In the $d=3$ case, it leads to a dilute Coulomb gas of
instantons, as shown by Polyakov~\cite{dilute}
long ago, while the corresponding instantons in $d=4$
share this property \cite{OT} and can
possibly also enable the construction of a dilute
Coulomb gas \cite{so4inst}. This is our physical
justification for making a systematic study of these
Higgs models, and we concentrate on the
$d=2,3$ cases here. Thus both models under consideration
here are the familiar ones, namely
the Abelian Higgs model in $d=2$ and the Georgi--Glashow
model in $d=3$.
The instantons in these two, $1+1$ and $2+1$ dimensional
models, are the well
known topologically stable Nielsen-Olesen vortices \cite{NO}
and the 't Hooft--Polyakov
monopoles \cite{tHP}, respectively. It is therefore the
sphaleron solutions to these models that
are the remaining entities to be studied. In this connection,
it is true that the sphalerons
of the Abelian Higgs model were studied extensively long
ago \cite{BS,C}, but we repeat it here for the sake of completeness so
that both cases $d=2$ and $d=3$ be treated similarly,
and, because the presentation of the results in the literature can be refined.
In particular we clarify the situation
with respect to the question of periodic boundary conditions in the
spacelike coordinate used in the literature~\cite{BS,C}.
The sphaleron in the $d=3$ case has been studied recently~\cite{Copeland}.
We have carried out the analysis of the sphaleron solutions in both models
employing both the non-contractible loop (NCL) used by Manton~\cite{Manton} for
the Weinberg-Salam model, to which we refer in this paper as the
{\it geometric} loop construction, as well as the the finite energy path
method of Akiba et al~\cite{AKY}, to which we refer as {\it boundary} loop
construction.
In the study of the sphaleron solutions, a central role is played
by the Chern--Simons number, which in the Weinberg--Salam model is calculated
from the second Chern--Pontryagin (CP) density. In $d$ dimensional
Higgs models we consider, the natural
candidate for the latter is the dimensionally reduced CP density
of the YM field on ${\rm I \hspace{-0.9mm} R}^d \times S^{4p-d}$, where $4p>d$~\cite{desc}.
The $d$ dimensional
model is decided by $p$ such that the action density is bounded from below
by dimensionally reduced the CP density in question.
The simplest such model corresponds to
the case where $4p$ is the smallest number greater than $d$. Both the Abelian
Higgs model and the Georgi--Glashow models considered in this work are the
simplest models whose CP densities are arrived at by the corresponding
dimensional descent of the {\it second} CP density of the $SU(2)$ YM field.
The analysis of the Abelian Higgs system in 2
dimensions is presented in Section 2. In
Subsection 2.1 and 2.3 respectively, the {\it geometric} and {\it boundary}
loop constructions are presented, while in Subsection 2.2 a discussion of our
constructions is contrasted with what is usually given in the literature, and
the differences are commented on.
The analysis of the Georgi-Glashow model in 3 dimensions is presented in
Section 3. In Subsection 3.1 the {\it boundary} loop construction is presented.
The {\it geometric} loop construction is presented in Subsection 3.2 . Since
the sphaleron solution of this model turns out to be effectively the solution
of an Abelian Higgs model, a family of arbitrary vorticity $N$ sphalerons are
contructed in Subsection 3.2 . Subsection 3.3 is devoted to a brief discussion
of the magnetic and electric properties of this sphaleron and the
configurations on the geometric loop. In section 4, we
give a summary of our results.
\section{$d=2$: Sphalerons in the Abelian Higgs model}
Our objective in this section is to construct noncontractible loops (NCL)
between two topologically neighbouring vacua which feature the sphaleron at
the top of the energy barrier. Our constructions run exactly parallel to those
of Manton \cite{Manton} and Akiba {\em et.al} \cite{AKY} respectively.
This differs from the analysis of Bochkarev and Shaposhnikov \cite{BS} in
that the latter employ periodicity in the space variable to construct the
NCL.
The $SO(2)$ Abelian Higgs model in $d=2$ spacetime dimensions is
given by the Euclidean Lagrange density
\begin{equation}
{\cal L}=\frac14F_{\mu\nu}^2+\frac12|D_{\mu}\varphi|^2+
\frac{\lambda_0}{32}(1-|\varphi|^2)^2
\label{a1}
\end{equation}
with $\varphi=\phi^1+i\phi^2$,
$D_{\mu}\varphi=\partial_{\mu}\varphi+iA_{\mu}\varphi$, i.e.\ we used the
identity $SO(2)=U(1)$ to write the Higgs dublett $\phi^i$ as one complex
field.
Adding the inequalities
\begin{equation}
\left(F_{\mu\nu}-\frac{\sqrt{\lambda_0}}{4}
\epsilon_{\mu\nu}(1-|\varphi|^2)\right)^2\ge 0, \qquad
\left|D_{\mu}\varphi-i\epsilon_{\mu\nu}D_{\nu}\varphi\right|^2\ge 0,
\label{ac1}
\end{equation}
the cross terms of the squares yield a lower bound for the
Euclidean action, ${\cal L}\ge \varrho$ which becomes a topological lower bound,
i.e.\ $\varrho$ can be written as total divergence
\begin{equation}
\varrho = \partial_{\mu}\Omega_{\mu}, \quad
\Omega_{\mu}=\frac14\epsilon_{\mu\nu}\left(A_{\nu}+
i\varphi^*D_{\mu}\varphi\right)
\label{a3}
\end{equation}
if $\lambda_0=1$.
$\Omega_{\mu}$ is the Chern--Simons form of this model.
It consists of a gauge
dependent and a gauge independent part which is a typical feature of
Higgs theories in even spacetime dimensions \cite{desc}.
The topological lower bound (which can be generalised to $\lambda_0\neq 1$)
ensures the stability of the instantons of the model, the Abelian Higgs
vortex, which can be found using the radially symmetric ansatz
\begin{equation}
\varphi=h(r)e^{-iN\theta} ,\qquad A_{\mu}=\frac{a(r)-N}{r} \ ,
\epsilon_{\mu\nu}\hat{x}_{\nu}
\label{aa1}
\end{equation}
with $r^2 = x^2_0 +x^2_1=t^2+x^2$,
and integrating the Euler--Lagrange equations of the resulting radial
subsystem Lagrangian
\begin{equation}
L_0=\pi \left\{ {1\over r}a'^2 +r\left(h'^2 + \frac{a^2 h^2}{r^2}
\right) +\frac{\lambda_0}{16} (1-h^2)^2 \right\} \ .
\label{abhiggsvortex}
\end{equation}
The corresponding instanton solutions are self--dual provided $\lambda_0=1$,
then they saturate the
inequality ${\cal L}\ge\partial_{\mu}\Omega_{\mu}$
and hence are characterised by integer Chern--Simons number (also
called Chern--Pontryagin charge in the context of instanton physics)
\begin{equation}
{\cal N}_{CS}= \frac{2}{\pi}\int_{-\infty}^{\infty}dt\int_{-\infty}^{\infty}dx\:
\partial_{\mu}\Omega_{\mu}=
\frac{2}{\pi}\int_{{\rm I \hspace{-0.9mm} R}^2}\partial_{\mu}\Omega_{\mu}d^2x=
\frac{2}{\pi}\lim_{r\rightarrow\infty}\int_{S^1}\Omega_{\mu}dS_{\mu}.
\label{aa3}
\end{equation}
This can be arrived at by subjecting the second Chern-Pontryagin class of the
$SU(2)$ YM field on ${\rm I \hspace{-0.9mm} R}^2 \times S^2$ to dimensional descent~\cite{desc}.
The Chern--Simons number depends only
on the behaviour of the instanton at infinity, reflecting the topological
properties of the mapping
\begin{equation}
\varphi_{inst}^{\infty}:S^1_{spacetime}\rightarrow S^1_{Higgs}.
\label{ab1}
\end{equation}
In contrast to the instanton, the sphaleron is a {\em ``static''}
object, i.e.\ it does not depend on the Euclidean time. Nontheless, it is an
{\em Euclidean} spacetime
object closely related to the instanton as it is the extremum
of the energy functional of a set of
static configurations connecting distinct vacua in a topologically nontrivial
way, i.e.\ a {\em noncontractible loop} (NCL).
The NCL, parametrised in
terms of Euclidean time, is an instanton--like object
and the loop as a whole having integer
topological number.
The Chern-Simons number ${\cal N}_{CS}$ is also parametrized in terms of the Euclidean
time,
\begin{equation}
{\cal N}_{CS}(t_0)= \frac{2}{\pi}\int_{-\infty}^{t_0}dt\int_{-\infty}^{\infty}dx\:
\partial_{\mu}\Omega_{\mu},
\label{Ncstime}
\end{equation}
hence the loop can be parametrizes in terms of ${\cal N}_{CS}$, the latter taking
on all values between $0$ and $1$.
To construct and study the sphaleron, we start from the static energy which in
the temporal gauge ($A_0=0$, $A:=A_1$) is given by
\begin{equation}
{\cal E}[\varphi,A]=\int\left[\frac12|(\partial_x+iA)\varphi|^2+
\frac{1}{32}(1-|\varphi|^2)^2\right]dx \ .
\label{a4}
\end{equation}
Choosing the particular ansatz $\varphi=\phi\in{\rm I \hspace{-0.9mm} R}$, $A=0$, this
reduces to the well--known real $\varphi^4$ energy functional
\begin{equation}
{\cal E}_{sph}[\phi]=\int\left[\frac12(\phi')^2
+\frac{1}{32}(1-\phi^2)^2\right]dx
\label{aaa4}
\end{equation}
which is minimized by the kink/antikink solutions
\begin{equation}
\phi_{\pm}(x)=\pm\tanh\left(\frac{x-x_0}{4}\right).
\label{ad1}
\end{equation}
The $\phi^4$ kink on its own is a stable soliton, but it becomes instable
as soon as it is embedded in complex isospace which can be shown explicitly
by investigating the fluctuation spectrum \cite{FH}.
\subsection{The geometrical loop construction}
To show that
$(\varphi,A)_{sph}=(\phi_{\pm},0)$ are sphaleron configurations
of the Abelian Higgs model, we construct the corresponding NCL
of finite energy configurations connecting two vacua through the
sphaleron in a topologically nontrivial way.
{\em Vacua} are static configurations with zero energy, i.e.\ they are given by
$\varphi=g$, $A=ig^{-1}\partial_xg$ with $g\in U(1)$.
We use the remaining gauge
freedom of the theory to fix the vacuum to $(\varphi,A)_{vac}=(1,0)$ and
comment on the choice of gauge later on.
Following the geometrical loop construction of Manton \cite{Manton}, we
consider first the topological properties of the Higgs field at infinity.
One dimensional space at infinity shrinks to the discrete set
$({\rm I \hspace{-0.9mm} R})^{\infty}= S^0_{space}=\{\pm1\}$. Hence one has to distinguish
two mappings $\varphi^{+\infty}$, $\varphi^{-\infty}$ instead of one mapping
$\varphi^{\infty}$ which depends on continous angular coordinates in higher
space dimensions. To construct a geometrical NCL \cite{Manton},
we introduce a single loop
parameter $\tau\in S^1_{loop}$ as additional degree of freedom such that
\begin{equation}
\varphi^{+\infty}:S^1_{loop} \rightarrow S^1_{Higgs}
\label{ae1}
\end{equation}
is a topologically
nontrivial mapping which we choose to be
\begin{equation}
\varphi^{+\infty}:\tau\mapsto e^{2i\tau},
\label{af1}
\end{equation}
whereas $\varphi^{-\infty}\equiv1$ is chosen to be topologically trivial.
The finite energy condition then also fixes the gauge field at infinity as
the covariant space derivative in (\ref{a4}) has to vanish, hence
\begin{equation}
A^{\pm\infty}:=
A(r\rightarrow\pm\infty)
=-i(\varphi^{\pm\infty})^{-1}\partial_x\varphi^{\pm\infty}=0.
\label{ag1}
\end{equation}
The general Manton loop ansatz constructed using the topological ingredients
$\varphi^{+\infty}$, $A^{\pm\infty}$ is now given by
\begin{equation}
\bar{\varphi}=(1-h(x))\psi+h(x)\varphi^{+\infty},\quad A=fA^{\pm\infty}=0
\label{ah1}
\end{equation}
where
$\psi$ has to be chosen such that the loop starts and ends in the vacuum and
reaches the sphaleron for $\tau=\frac{\pi}{2}$, hence $\bar{\varphi}|_{\tau=0}=
\bar{\varphi}|_{\tau=\pi}=1$,
$\bar{\varphi}|_{\tau=\frac{\pi}{2}}=h\varphi^{+\infty}|_{\tau=\frac{\pi}{2}}=-h$.
In the $SO(2)$ model,
\begin{equation}
\psi=\cos^2\tau+i\sin\tau\cos\tau
\label{ai1}
\end{equation}
is a convenient choice. Moreover,
$\bar{\varphi}(x\rightarrow\pm\infty)=\varphi^{\pm\infty}$ requires
$h(x\rightarrow\pm\infty)=\pm 1$.
The loop is then given by
\begin{equation}
\bar{\varphi}(\tau,x)=e^{i\tau}[\cos\tau+ih(x)\sin\tau],\qquad \bar{A}\equiv 0.
\label{a5}
\end{equation}
Inserting the ansatz (\ref{a5}) into the static energy functional (\ref{a4})
yields
\begin{equation}
{\cal E}[\bar{\varphi},\bar{A}]=\bar{{\cal E}}_{\tau}[h]=\sin^2\tau \int\left[\frac12(h')^2+
\frac{1}{32}\sin^2\tau(1-h^2)^2\right]dx
\label{a6}
\end{equation}
which for $\tau=\frac{\pi}{2}$ reduces to the $\varphi^4$ model,
$\bar{{\cal E}}_{\tau=\frac{\pi}{2}}={\cal E}_{sphal}=\frac{1}{3}$.
For any fixed value of $\tau\in[0,\pi]$, $\bar{{\cal E}}_{\tau}[h]$ is
{\em minimised} by
\begin{equation}
h_{\tau}=\tanh\left(\sin\tau\frac{x-x_0}{4}\right).
\label{aa6}
\end{equation}
The energy along the resulting minimal energy loop is
\begin{equation}
\bar{{\cal E}}(\tau)=\frac13\sin^3\tau
\label{ab6}
\end{equation}
which has a {\em maximum} for
$\frac{\pi}{2}$.
This minimax procedure therefore shows that $(\varphi,A)_{sph}=
(\phi_{-},0)$ is a sphaleron of the Abelian Higgs model.
To construct the NCL for the kink--type sphaleron $(\varphi,A)_{sph}=
(\phi_{+},0)$, one simply has to use
\begin{equation}
\varphi^{-\infty}:\tau\mapsto e^{2i\tau}
\label{ac6}
\end{equation}
as the topologically nontrivial mapping.
To calculate the increase of the Chern--Simons number
along the NCL, one has to treat the loop parameter as a
(Euclidean) time dependent quantity
$\tau=\tau(t)$ with $\tau(t=-\infty)=0$, $\tau(t=\infty)=\pi$. Inserting
the loop ansatz (\ref{a5}) into (\ref{Ncstime}),
one can split the double integral into
a space ``volume'' and a ``surface'' integral, resulting in
\begin{equation}
{\cal N}_{CS}(t_0) = \frac{2}{\pi}\left\{\int_{-\infty}^{\infty}
\Omega_0\big|^{t=t_0}_{t=-\infty}dx +
\int_0^{t_0}\Omega_1\big|_{x=-\infty}^{x=+\infty}dt\right\}
=\frac{\tau(t_0)}{\pi}+\frac{1}{2\pi}\sin 2\tau(t_0).
\label{a8}
\end{equation}
As expected, one finds ${\cal N}_{CS}=\frac12$ when the loop reaches the sphaleron,
whereas the entire loop has ${\cal N}_{CS}=1$.
\subsection{Gauging and nonperiodic boundary conditions}
The vacuum convention chosen above can be changed by {\em gauging the
geometrical loop},
\begin{equation}
\bar{\varphi}\rightarrow \check{\varphi}=g\bar{\varphi},\qquad \bar{A}=0\rightarrow \check{A}=
ig^*\partial_xg.
\label{aa8}
\end{equation}
One could, e.g., require $\check{\varphi}=g\bar{\varphi}\rightarrow 1$
for $x\rightarrow -\infty$ {\em and} $x \rightarrow +\infty$, choosing the
gauge \cite{FH}
\begin{equation}
g=e^{-i\tau\Lambda(x)},\qquad\Lambda(x)=\frac{2}{\pi}\arctan
\left(\frac{x}{\alpha}\right)+1
\label{ac8}
\end{equation}
($\alpha\neq 0$ arbitrary)
which yields $A(x)=\frac{2\tau}{\pi}
\frac{\alpha}{\alpha^2+x^2}$. Concerning its $\tau$--dependence, $g$ can be
considered either as a set of static transformations parametrised by $\tau$
and applied to each corresponding static configuration along the loop
separately, or as one time dependent gauge transformation applied to the
$\tau(t)$ time dependent loop as a whole. In the latter case, the temporal
gauge condition is violated, $A_0\equiv 0 \rightarrow \dot{\tau}\Lambda$.
In both cases, one finds for the gauge transformed Chern--Simons form
$\Omega_0\rightarrow\Omega_0+\frac14\tau\Lambda'$,
$\Omega_1\rightarrow\Omega_1-\frac14\dot{\tau}\Lambda$. Inserting this into
eq.\ (\ref{a8}), we find explicitly that the Chern--Simons number is gauge
invariant, i.e.\ also the gauged Manton loop $(\check{\varphi},\check{A})$
starts at a vacuum with
${\cal N}_{CS}(\tau=0)=0$, reaches the sphaleron at
${\cal N}_{CS}\left(\tau=\frac{\pi}{2}\right)=\frac12$ and ends a vacuum
with ${\cal N}_{CS}(\tau=\pi)=1$. This is not surprising, as $\varrho=
\partial_{\mu}\Omega_{\mu}$ is a gauge invariant quantity.
The main new feature of the gauged geometrical
loop is the representation of the
vacuum states between which the loop interpolates. Generalizing the gauged
geometrical loop $(\check{\varphi},\check{A})$ to a larger range of the loop
parameter $\tau\in{\rm I \hspace{-0.9mm} R}$, the loop reaches a vacuum state
for any $\tau=n\pi$, $n\in{\rm Z \hspace{-1.2mm} Z}$, now given by
\begin{equation}
\check{\varphi}^{(n)}=\left(-\frac{\alpha-ix}{\alpha+ix}\right)^n
\qquad \check{A}^{(n)}=\frac{n\alpha}{\alpha^2+x^2}
\label{a9}
\end{equation}
These vacua can be labeled by a ``topological charge'' defined as
\begin{equation}
Q:=\frac{2}{\pi}\int\Omega_0dx \quad \Rightarrow \quad
Q\big|_{(\check{\varphi}^{(n)},\check{A}^{(n)})}=\frac{1}{2\pi}\int \check{A}^{(n)} dx = n.
\label{a10}
\end{equation}
In the literature \cite{BS,C} this is usually the only ``topological number''
discussed in the context of the $SO(2)$ Abelian Higgs sphaleron. It
should be stressed that it is different from the Chern--Simons number
which is the proper quantity to be ivestigated in the context of the
sphaleron NCL and its relation to the instanton. The topological
charge used above to label the vacua makes use of the gauge--variant
part of the Chern--Simons form. Only in even dimensions does the Chern--Simons
form exhibit a gauge--variant part~\cite{desc}.
In these dimensions one can distinguish between
``small'' and ``large'' gauge transformations, depending on whether
they leave the ``topological charge'' (defined as space volume integral
of the zero component of the Chern--Simons form) invariant or not.
Another confusion about the $SO(2)$ Abelian Higgs found in the
literature \cite{BS,C} concerns the use of periodic boundary conditions in the
space coordinates for the gauge fields, i.e.\ ``putting the fields on the
circle'' instead of using the non--periodic vacuum structure (\ref{a9})
\cite{FH}. The sphaleron, however, is definitely
never a periodic object in space, but always the $\varphi^4$ kink, a
contradiction which cannot be solved by simpy taking the limit of infinite
period.
The periodic solutions of $\varphi^4$ theory \cite{ManSam} and
Goldstone theory \cite{BrihTom} in one dimension can be more gainfully
interpreted as the {\em periodic instantons in Euclidean quantum mechanics}.
These
periodic solutions describe tunneling from thermally excited states, the
temperature being given by the inverse period. At some value of the period,
the periodic instantons reduce to the (time independent) sphaleron which in
the case of Euclidean quantum mechanics is just a constant solution. This
effect describes the phase transition between classical and quantum behaviour
which is presently under intense investigation \cite{perinst}. Since periodic
instantons (periodic in the Euclidean time, not in the spatial coordinate)
are also known to exist in the Abelian Higgs model \cite{M}, one should be
even more careful about the notion of periodicity in this model.
\subsection{AKY boundary loop construction}
A different technique for the construction of a NCL for the Abelian Higgs
sphaleron is
motivated by the {\em ``static minimal energy path''} construction by
Akiba, Kikuchi and Yanagida (AKY) \cite{AKY} in the Weinberg--Salam theory.
This NCL is constructed by minimizing a general spherically symmetric static
ansatz with parameter-dependent boundary conditions, i.e.\ not the loop
ansatz containing
the loop parameter, but the boundary conditions. We shall refer to
this type of construction as ``boundary loop''.
The spherically symmetric AKY ansatz in one space dimension (in temporal
gauge) is simple and only puts some restriction on the symmetry
properties of the parameter functions,
\begin{equation}
\tilde{\varphi}(x)=H(|x|)+i\hat{x}K(|x|), \qquad \tilde{A}(x)=f(x)
\label{a11}
\end{equation}
with $\hat{x}=\mbox{sgn}(x)$, i.e.\ $H(x)=H(|x|)$ is an even, $H(x)=\hat{x}H(|x|)$
an odd function of $x\in{\rm I \hspace{-0.9mm} R}$. Under a gauge transformation $g=e^{i\Lambda(x)}$,
$H$ and $K$ are rotated,
\begin{equation}
H\rightarrow H\cos\Lambda-K\sin\Lambda, \qquad
K\rightarrow K\cos\Lambda+H\sin\Lambda,
\label{aa11}
\end{equation}
while $f\rightarrow f-\Lambda'$,
i.e.\ one can gauge $f=0$ without loss of generality in the ansatz (\ref{a11}).
Inserting the ansatz into the static energy functional yields
\begin{equation}
{\cal E}[\tilde{\varphi},\tilde{A}]=
\tilde{{\cal E}}[H,K]=
\int\left[\frac12(H'+K')^2+\frac{1}{32}(1-H^2-K^2)^2\right]dx,
\label{a12}
\end{equation}
i.e. finite energy requires
\begin{equation}
H(x\rightarrow\infty)=\cos q, \qquad K(x\rightarrow
\infty)=\sin q.
\label{a12aa}
\end{equation}
Extrema of $\tilde{{\cal E}}[H,K]$ with these boundary conditions
(and $H$ even, $K$ odd) are found only for $q=0$ or $q=\pi$, yielding the
vacua $H=\pm 1$, $K=0$, and for $q=\frac{\pi}{2}$ with $H=0$,
$K(x)=\phi_+(x)=\tanh\left(\frac{x}{4}\right)$ which is the sphaleron
configuration.
The idea now is to construct a loop by increasing the boundary
condition parameter $q(t)$
as a time--dependent quantity from $q(t=-\infty)=0$ to $q(t=\infty)=\pi$.
Inserting the ansatz in the Chern--Simons functional (\ref{a8}) and taking
care of the boundary conditions for $H$, $K$ when evaluating
$\Omega_1\big|_{x=-\infty}^{x=+\infty}$ yields
\begin{equation}
{\cal N}_{CS}=\frac{1}{2\pi}\int(KH'-HK')dx+\frac{q}{\pi},
\label{aa12}
\end{equation}
i.e.\ the
sphaleron has ${\cal N}_{CS}=\frac12$. One can minimise the static energy functional
(\ref{a12}) for fixed Chern--Simons number by adding the Chern--Simons
functional with a Lagrange multiplier $\xi$ to the static energy functional,
leading to the Euler--Lagrange equations
\begin{equation}
H''+\xi K' + \frac18H(1-H^2-K^2) = 0, \qquad
K''-\xi H' + \frac18H(1-H^2-K^2) = 0
\label{a13}
\end{equation}
which are solved with the above boundary conditions by
\begin{equation}
H(x)=\cos q,\qquad K(x)=\sin q \tanh\left(\frac{\sin q}{4}x\right),
\qquad \xi=8\cos q.
\label{aa13}
\end{equation}
The energy
along this loop is found to be
\begin{equation}
\tilde{{\cal E}}(q)=\frac13\sin^3 q,
\label{ab13}
\end{equation}
and the Chern--Simons number
along the AKY loop for the Abelian Higgs model is
\begin{equation}
\tilde{{\cal N}}_{cs}(q)=\frac{q}{\pi}+\frac{1}{2\pi}\sin 2q.
\label{ac13}
\end{equation}
Therefore, the energy along the loop, $\tilde{{\cal E}}(\tilde{{\cal N}}_{cs})$ agrees
with the result (\ref{ab6},\ref{a8}) for the geometrical NCL. Hence there
is no advantage of lower energy along the boundary loop compared to
the geometrical loop, both provide equivalent
``static minimal energy paths'' \cite{AKY}.
\section{$d=3$: Sphalerons in the Georgi--Glashow model}
The model is described by the $SO(3)$ taking its values in the $SU(2)$
algebra with antihermitean generators $-\frac{i}{2}(\sigma_{\mu})=
-\frac{i}{2}\vec{\sigma}$ and a
Higgs triplet field $(\phi^{\mu})=\vec{\phi}$ which we write in antihermitean
isovector representation,
$\Phi=\vec{\phi}\cdot\left(-\frac{i}{2}\vec{\sigma}\right)$, in
$d=3$ spacetime
dimensions. The model is given by the Euclidean Lagrangian
\begin{eqnarray}
{\cal L} & = &
\mbox{tr}\left[-\frac14F_{\mu\nu}^2 - \frac12 (D_{\mu}\Phi)^2 + \frac{\lambda}{2}
\left(\Phi^2+\frac{\eta^2}{2}\right)^2\right]
\label{b1} \\
& = & \frac12\left[\frac14\vec{F}^2_{\mu\nu}+
\frac12\left(D_{\mu}\vec{\phi}\right)^2+\frac{\lambda}{8}\left(\vec{\phi}^2-
2\eta^2\right)^2\right]
\label{b1pol}
\end{eqnarray}
with $F_{\mu\nu}=\partial_{[\mu}A_{\nu]}+[A_{\mu},A_{\nu}]
=\vec{F}_{\mu\nu}\cdot\left(-\frac{i}{2}\vec{\sigma}\right)$ and
$D_{\mu}\Phi=\partial_{\mu}\Phi+[A_\mu,\Phi]\Rightarrow
D_{\mu}\vec{\phi}=\partial_{\mu}\vec{\phi}+\vec{\phi}\wedge\vec{A}_{\mu}$,
with $\mu=1,2,3$.
>From the second form of the Lagrangian \cite{dilute} it is easy to identify
the particle spectrum of the theory which consists of photons, heavy
charged bosons with mass $m_W=\sqrt{2}\eta$, and scalar neutral Higgs
particles with mass $m_H=\sqrt{2\lambda}\eta$.
The model was previously
exploited by Polyakov \cite{dilute} to calculate quark confinement effects,
using a dilute gas of instantons. The instanton of the model is the
finite size spherically symmetric
't Hooft--Polyakov monopole
\cite{tHP} in three dimensions which is
characterised by integer Chern--Simons number
\begin{equation}
{\cal N}_{CS}=
\frac{1}{\sqrt{2}\pi\eta}\int\partial_{\mu}\Omega_{\mu}d^3x\quad \mbox{with}
\quad
\Omega_{\rho}=\frac14\epsilon_{\rho\mu\nu}\mbox{tr}\left[\Phi F_{\mu\nu}\right].
\label{b1a}
\end{equation}
As the model is in an odd spacetime dimension, the Chern--Simons form
$\Omega_{\mu}$ in this model,
which is constructed from the Bogomol'nyi inequality
\begin{equation}
\mbox{tr}\left[F_{\mu\nu}-\epsilon_{\mu\nu\rho}D_{\rho}\Phi\right]^2>0\Rightarrow
{\cal L}>\partial_{\mu}\Omega_{\mu} \ ,
\label{b1b}
\end{equation}
has no gauge variant part \cite{desc}.
To discuss the sphalerons of the model \cite{Copeland},
we reduce the Langrangian (\ref{b1})
to the static Hamiltonian
\begin{equation}
{\cal H}=\mbox{tr}\left[-\frac14F_{ij}^2 - \frac12 (D_{i}\Phi)^2 +\lambda
\left(\Phi^2+\frac{\eta^2}{2}\right)^2\right]
\label{b2}
\end{equation}
with $i,j,\ldots=1,2$.
\subsection{Sphaleron and boundary loop construction}
\label{so3aky}
Motivated by the AKY technique in the Weinberg--Salam model \cite{AKY},
we try the general static spherically symmetric ansatz for the Higgs field
to find the sphaleron and a corresponding NCL simultanously,
\begin{equation}
\Phi=i\frac{\eta}{\sqrt{2}}\left[H(r)\sigma_3+K(r)\hat{x}_i\sigma_i\right] \ ,
\label{b3}
\end{equation}
with $r^2=x^2_1+x^2_2$.
The $SO(2)$ gauge transformation $\Phi\rightarrow g^{-1}\Phi g$ with
\begin{equation}
g_{\Lambda}
=\exp\left\{\frac{i}{2}\epsilon_{ij} \hat{x}_i\sigma_j\Lambda(r)\right\}
\label{gGauge}
\end{equation}
rotates the parameter functions,
\begin{equation}
H\rightarrow H\cos\Lambda-K\sin\Lambda,\qquad
K\rightarrow K\cos\Lambda+H\sin\Lambda.
\label{b3a}
\end{equation}
This motivates the choice of a spatial spherically symmetric ansatz
\begin{equation}
A_i=\frac{f_A+1}{r}\epsilon_{ik}\hat{x}_k\left(-\frac{i}{2}\sigma_3\right) +
\frac{f_B}{r}\epsilon_{ik}\hat{x}_k\left(-\frac{i}{2}\hat{x}_i\sigma_i\right)+
\frac{f_C}{r}\hat{x}_i \left(-\frac{i}{2}\epsilon_{kl}\hat{x}_k\sigma_l\right)
\label{b4}
\end{equation}
for the gauge field
which transforms to an ansatz of the same type under
$A_i\rightarrow g^{-1}A_ig+g^{-1}\partial_{\mu}g$, in particular,
\begin{equation}
f_A\rightarrow f_A\cos\Lambda-f_B\sin\Lambda,\quad
f_B\rightarrow f_B\cos\Lambda+f_A\sin\Lambda,\quad
f_C\rightarrow f_C-r\Lambda'.
\label{b4a}
\end{equation}
The requirement of regularity at the origin implies
\begin{equation}
f_A(0)=-1,\quad f_B(0)=0, \quad f_C(0)=0, \quad H(0)=const, \quad K(0)=0,
\label{zero}
\end{equation}
Also the gauge
transformation has to be regular at the origin, $\Lambda(0)=0$.
In this gauge, inserting the ansatz (\ref{b3},\ref{b4}) reduces
the static Hamiltonian to the following radial subsytem Hamiltonian:
\begin{eqnarray}
\tilde{H}_0& = &
2\pi\Bigg\{
\frac{1}{4r}\left[\left(f_A^{\prime}-\frac{f_Bf_C}{r}\right)^2+
\left(f_B^{\prime}+\frac{f_Af_C}{r}\right)\right]
\nonumber \\
& & {} +\frac{\eta^2}{2}\left[r\left(H^{\prime}-\frac{Kf_C}{r}\right)
+r\left(K^{\prime}+\frac{Hf_C}{r}\right)+\frac{1}{r}(Kf_A-Hf_B)^2\right]
\nonumber \\
& & {} +\frac{\lambda\eta^4}{4} r(1-H^2-K^2)^2\Bigg\}
\label{b5general}
\end{eqnarray}
The Euler--Lagrange equations are
\begin{eqnarray}
\left[r\left(H^{\prime}-\frac{Kf_C}{r}\right)\right]^{\prime}
& = &
f_C\left(K^{\prime}+\frac{Hf_C}{r}\right)
-\frac{1}{r}f_B(Kf_A-Hf_B)
-\lambda\eta^2rH(1-H^2-K^2)
\label{Eins}
\\
\left[r\left(K^{\prime}+\frac{Hf_C}{r}\right)\right]^{\prime}
& = &
-f_C\left(H^{\prime}-\frac{Kf_C}{r}\right)
+\frac{1}{r}f_A(Kf_A-Hf_B)
-\lambda\eta^2rK(1-H^2-K^2) \nonumber \\
\label{Zwei}
\\
\left[\frac{1}{r}\left(f_A^{\prime}-\frac{f_Bf_C}{r}\right)\right]^{\prime}
& = &
\frac{1}{r^2}f_C\left(f_B^{\prime}+\frac{f_Af_C}{r}\right)
+\frac{2\eta^2}{r}K(Kf_A-Hf_B)
\label{Drei}
\\
\left[\frac{1}{r}\left(f_B^{\prime}+\frac{f_Bf_C}{r}\right)\right]^{\prime}
& = &
-\frac{1}{r^2}f_C\left(f_A^{\prime}-\frac{f_Bf_C}{r}\right)
-\frac{2\eta^2}{r}H(Kf_A-Hf_B)
\label{Vier}
\end{eqnarray}
and
\begin{equation}
0 = f_A\left(f_B^{\prime}+\frac{f_Bf_C}{r}\right)
-f_B\left(f_A^{\prime}-\frac{f_Bf_C}{r}\right)
+2\eta^2r^2\left[H\left(K^{\prime}+\frac{Hf_C}{r}\right)
-K\left(H^{\prime}-\frac{Kf_C}{r}\right)\right]
\label{Fuenf}
\end{equation}
Note that the ansatz (\ref{b3},\ref{b4}) being radially symmetric, eqs.\
(\ref{Eins}--\ref{Fuenf}) are guaranteed
consistent with the equations which are obtained by varying the original
Hamiltonian (\ref{b2}) and inserting the ansatz (\ref{b3},\ref{b4}) afterwards.
This is in contrast with the situation in the Weinberg--Salam model where the
corresponding ansatz is not spherically symmetric and hence the
consistency of the ansatz there must be checked.
As the fifth equation (\ref{Fuenf}) is first--order, it is obviously not a
dynamical equation, but a constraint on the system which is related to
a gauge transformation in the space of the parameter functions
$(H,K,f_A,f_B,f_C)$ \cite{constraint}.
This shows that this set of parameter functions has one
redundant degree of freedom which is already clear from the fact that we
kept the gauge freedom (\ref{gGauge},\ref{b3a},\ref{b4a}) in deriving
the reduced Hamiltonian (\ref{b5general}).
Indeed, the gauge transformation generated by eq.\ (\ref{Fuenf}) is equivalent
to the transformations (\ref{b3a},\ref{b4a}) derived from the
$SO(3)$ gauge transformation $g$ of the original fields $(\Phi,A_{i})$
in (\ref{gGauge}).
The gauge transformation (\ref{b3a},\ref{b4a}) can now be used to fix
$f_C\equiv 0$ without loss of generality in the ansatz (\ref{b3}). This
choice corresponds to ``radial gauge'' $\hat{x}_iA_i=0$. In this gauge,
the Hamiltonian (\ref{b5general}) simplifies to
\begin{equation}
H_0=2\pi
\left\{\frac{1}{4r}(f_A^{\prime 2}+f_B^{\prime 2}) +
\frac{\eta^2}{2}\left[r(H^{\prime 2}+K^{\prime 2})+\frac{1}{r}(Kf_A-Hf_B)^2
\right]
+\frac{\lambda\eta^4}{4}r(1-H^2-K^2)^2\right\}
\label{b5}
\end{equation}
For $f_B\equiv 0$ and $H\equiv 0$ this reduces to the radial subsystem
Euclidean Lagrangian of the Abelian Higgs model in two spacetime dimensions
(\ref{abhiggsvortex}) supporting the (topologically stable)
Abelian Higgs vortex. The embedding of this two dimensional
$SO(2)$--Higgs vortex into the $SO(3)$--Higgs theory as a
static object in three spactime dimensions, $(\Phi,A_i)_{sph}$
yields the sphaleron of the three dimensional model.
It is interesting to remark that the $SO(3)$ gauge field of this sphaleron
tends to one half times a pure gauge at spatial infinity, $r\rightarrow\infty$:
\begin{equation}
(A_i)_{sph} \sim
\frac{1}{r}\epsilon_{ik}\hat{x}_k\left(-\frac{i}{2}\sigma_3\right)
=\frac12 g_{\pi}^{-1}\partial_ig_{\pi}, \qquad
g_{\pi}=\exp\left\{\frac{i}{2}\epsilon_{ij} \hat{x}_i\sigma_j\pi\right\},
\label{halfpuregauge}
\end{equation}
a property which the $SO(3)$ Higgs sphalerons shares with the instanton
of the same model \cite{dilute}. This is not surprising if we consider
that sphalerons and instantons are related objects.
Requiring finite energy
\begin{equation}
{\cal E}=\int{\cal H} d^2x=\int H_0 dr<\infty
\label{energy}
\end{equation}
fixes the behaviour of the remaining parameter functions $(H,K,f_A,f_B)$
at spatial infinity to
\begin{eqnarray}
H(r\rightarrow\infty)=\cos q & & f_A(r\rightarrow\infty)=-\alpha(q)\cos q
\nonumber \\
K(r\rightarrow\infty)=\sin q & &
f_B(r\rightarrow\infty)=-\alpha(q)\sin q
\label{infty}
\end{eqnarray}
with $q\in[0,\pi]$.
The relation between $q$ and $\alpha=\alpha(q)$
can be determined from a more careful
analysis of the asymptotic behaviour of the parameter functions
$(H,K,f_A,f_B)$, using the Euler--Lagrange equations of $H_0$ which are
\begin{eqnarray}
\left(\frac{f_A^{\prime}}{r}\right)^{\prime}
& = & \frac{2\eta^2}{r}K(Kf_A-Hf_B)
\label{eins}
\\
\left(\frac{f_B^{\prime}}{r}\right)^{\prime}
& = & -\frac{2\eta^2}{r}H(Kf_A-Hf_B)
\label{zwei}
\\
(r H^{\prime})^{\prime}
& = & -\frac{f_B}{r}f_B(Kf_A-Hf_B)-\lambda\eta^2 r H(1-H^2-K^2)
\label{drei}
\\
(r K^{\prime})^{\prime}
& = & \frac{f_A}{r}(Kf_A-Hf_B)-\lambda\eta^2 r K(1-H^2-K^2)
\label{vier}
\end{eqnarray}
Of course, these equations agree with (\ref{Eins}--\ref{Vier}), setting
$f_C=0$. The fifth equation (\ref{Fuenf}) reduces to
\begin{equation}
f_A\left(\frac{f_B^{\prime}}{r}\right)-
f_B\left(\frac{f_A^{\prime}}{r}\right)+
2\eta^2\left[H(r K^{\prime})-K(r H^{\prime})\right] = 0
\label{fuenf}
\end{equation}
which turns out to be an integration constant of the other four equations
(\ref{eins}--\ref{vier}).
The asymptotic analysis then fixes $\alpha(q)=\cos q$,
and one obtains the following asymptotic behaviour in the region $m_W r \gg1$:
\begin{eqnarray}
H(r) & \sim & +\cos q - d_H\cos q\cdot (m_Wr)^{-\frac12}
e^{-m_H r} - d_W \cos q \sin q \cdot (m_Wr)^{-\frac32}e^{-m_W r}
\label{asymptH} \\
K(r) & \sim & +\sin q - d_H \sin q \cdot (m_Wr)^{-\frac12}
e^{-m_H r}+ d_W \cos^2 q \sin q \cdot (m_Wr)^{-\frac32}
e^{-m_W r}
\label{asymptK} \\
f_A(r) & \sim & - \cos^2 q + D_W\sin q \cdot (m_Wr)^{\frac12}e^{-m_W r}
\label{asymptfa} \\
f_B(r) & \sim & - \cos q \sin q -
D_W\cos q \cdot (m_Wr)^{\frac12}e^{-m_W r}
\label{asymptfb}
\end{eqnarray}
where $D_W$, $d_H$, $d_W$ are constants which can be determined numerically
by solving eqs.\ (\ref{eins}--\ref{vier}).
It is easy to check that $m_W r =:\rho$ is the convenient dimensionless radial
variable in this model which is also used in all numerical calculations.
Solving eqs.\ (\ref{eins}--\ref{vier}) in the $m_W r\ll 1$ region yields
\begin{eqnarray}
H(r) & \sim & c_H+\frac14c_H(c_H^2-1)\cdot (m_Hr)^2 \label{smallH}\\
K(r) & \sim & c_K\cdot (m_Wr) \label{smallK} \\
f_A(r) & \sim & -1 + c_A \cdot (m_Wr)^2 \label{smallfA} \\
f_B(r) & \sim & c_B \cdot (m_Wr)^3 \label{smallfB}
\end{eqnarray}
with $c_A$, $c_B$, $c_K$ constants which again have to be determined from
the numerical integration of the
equations of motion, and $c_H=3\frac{c_B}{c_K}$.
Solutions of eqs.\ (\ref{eins}--\ref{vier}) can only be found for three
particular values of $q$, $q\in [0,\pi]$.
First, $q=0$ and $q=\pi$ allow the {\em vacuum configurations}
$H\equiv\pm 1$, $K\equiv 0$, $f_A\equiv -1$, $f_B\equiv 0$ with zero energy,
resulting in the vacuum fields
\begin{equation}
(\Phi,A_i)_{\pm}=\left(\pm i\frac{\eta}{\sqrt{2}}\sigma_3,0\right)
\label{vacua}
\end{equation}
Second for
$q=\frac{\pi}{2}$ the behaviour of the parameter functions at infinity
(\ref{asymptH}--\ref{asymptfb}) and at the origin (\ref{smallH}--\ref{smallfB})
allow to set $f_B\equiv 0$ and $H\equiv 0$ such that the $SO(2)$ symmetric
Abelian Higgs vortex configuration is an element of the subspace of $SO(3)$
Higgs configurations given by the ansatz (\ref{b3},\ref{b4}) and the boundary
conditions (\ref{zero},\ref{infty}) in the radial $f_C\equiv0$ gauge.
This is the sphaleron solution.
To construct a loop from the subspace of $SO(3)$ Higgs configurations
discussed above, we have to specify a one parameter subset
$(\tilde{\Phi},\tilde{A}_i)_q$ parametrised by the boundary constant $q$.
Three points of the loop are already fixed: The initial and final
vacua $(\tilde{\Phi},\tilde{A}_i)_{q=0,\pi}=(\Phi,A_i)_{\pm}$ and the sphaleron
$(\tilde{\Phi},\tilde{A}_i)_{q=\frac{\pi}{2}}=(\Phi,A_i)_{sph}$ given by the parameter
functions $(f_A,K)_{sph}$. We now have to fix the parameter functions
$(H,K,f_A,f_B)$ for all remaining values of $q$ such that $q$ becomes the
loop parameter.
Instead of $q$, it is convenient to choose the Chern--Simons
number as parameter along the loop. Therefore, we next
calculate the Chern-Simons functional for
the set of configurations (\ref{b3},\ref{b4}) with boundary behaviour
(\ref{asymptH}--\ref{asymptfb}) at infinity and (\ref{smallH}--\ref{smallfB})
at the origin, treating $q=q(t)$ as time dependent parameter such that a loop
parametrised by $q$ starts at an initial vacuum with $q(t=-\infty)=0$ and
ends at a final vacuum with $q(t=+\infty)=\pi$, whereas the sphaleron is
reached at some time $t_{sph}$, $q(t_{sph})=\frac{\pi}{2}$.
Inserting the ansatz (in radial gauge $f_C\equiv 0$)
into the Chern-Simons number functional
\begin{equation}
{\cal N}_{CS}(t_0) = \frac{1}{\sqrt{2}\pi\eta}\left\{\int_{{\rm I \hspace{-0.9mm} R}^2}
\Omega_0\big|^{t=t_0}_{t=-\infty}d^2x + \int_{-\infty}^{t_0} dt
\lim_{r\rightarrow\infty} \int_{S^1(r)}\Omega_idS_i\right\}
\label{bCS}
\end{equation}
we obtain
\begin{equation}
{\cal N}_{CS}(q)=-\frac12\int\left(Hf_A^{\prime}+Kf_B^{\prime}\right)dr
+\frac12(1-\cos q),
\label{so3ncs}
\end{equation}
with $q=q(t_0)$. This yields ${\cal N}_{CS}(q=0)=0$ and ${\cal N}_{CS}(q=\pi)=1$, hence
the Chern--Simons number increases by unity between the initial and the final
vacuum. For the sphaleron, the surface integral in (\ref{bCS}) yields
${\cal N}_{CS}\left(q=\frac{\pi}{2}\right)=\frac12$.
To complete the loop construction, we minimise the static energy (\ref{energy})
for fixed value of the Chern--Simons number ${\cal N}_{CS}$. This is done by
adding the integrand in (\ref{so3ncs}),
\begin{equation}
N_0=\frac12\left(Hf_A^{\prime}+Kf_B^{\prime}\right),
\label{lm}
\end{equation}
multiplied by a Lagrange multiplier $\xi$, to the reduced Hamiltonian
$H_0$ in (\ref{b5}), and minimising
\begin{equation}
F_0:=H_0+\xi N_0.
\label{withmult}
\end{equation}
The corresponding
Euler--Lagrange equations are
\begin{eqnarray}
\left(\frac{f_A^{\prime}}{r}\right)^{\prime}
& = & \frac{2\eta^2}{r}K(Kf_A-Hf_B) - \xi H^{\prime}
\label{leins}
\\
\left(\frac{f_B^{\prime}}{r}\right)^{\prime}
& = & -\frac{2\eta^2}{r}H(Kf_A-Hf_B) - \xi K^{\prime}
\label{lzwei}
\\
(r H^{\prime})^{\prime}
& = & -\frac{f_B}{r}f(Kf_A-Hf_B)-\lambda\eta^2 r H(1-H^2-K^2)
+\frac{\xi}{2\eta^2}f_A^{\prime}
\label{ldrei}
\\
(r K^{\prime})^{\prime}
& = & \frac{f_A}{r}(Kf_A-Hf_B)-\lambda\eta^2 r K(1-H^2-K^2)
+\frac{\xi}{2\eta^2}f_B^{\prime}.
\label{lvier}
\end{eqnarray}
The Lagrange multiplier changes the behaviour of the solutions at infinity
to
\begin{eqnarray}
H(r) & \sim & +\cos q - d_H\cos q\cdot (m_Wr)^{-\frac12}
e^{-\sqrt{m_H^2-\frac{\xi^2}{m_W^2}} r}
- \xi d_W \sin q \cdot (m_Wr)^{\frac12}
e^{-\sqrt{m_W^2-\frac{\xi^2}{m_W^2}} r}
\label{lasymptH} \\
K(r) & \sim & +\sin q - d_H \sin q \cdot (m_Wr)^{-\frac12}
e^{-\sqrt{m_H^2-\frac{\xi^2}{m_W^2}} r}+
\xi d_W \cos q \cdot (m_Wr)^{\frac12}
e^{-\sqrt{m_W^2-\frac{\xi^2}{m_W^2}} r}
\label{lasymptK} \\
f_A(r) & \sim & - \alpha(q)\cos q + \xi D_H \cos q \cdot (m_Wr)^{\frac12}
e^{-\sqrt{m_H^2-\frac{\xi^2}{m_W^2}} r} + D_W \sin q \cdot (m_Wr)^{\frac12}
e^{-\sqrt{m_W^2-\frac{\xi^2}{m_W^2}} r} \nonumber
\\ \label{lasymptfa} \\
f_B(r) & \sim & - \alpha(q) \sin q + \xi D_H \sin q \cdot (m_Wr)^{\frac12}
e^{-\sqrt{m_H^2-\frac{\xi^2}{m_W^2}} r} - D_W \cos q \cdot (m_Wr)^{\frac12}
e^{-\sqrt{m_W^2-\frac{\xi^2}{m_W^2}} r}, \nonumber
\\ \label{lasymptfb}
\end{eqnarray}
and requiring exponential decay restricts the range of the Lagrange
multiplier to
\begin{equation}
|\xi|<\min\{m_W^2,m_Wm_H\}.
\label{lmu}
\end{equation}
The asymptotic analysis no longer forces $\alpha(q)=\cos q$, and
$\alpha$ along the loop is known only for the vacua and the sphaleron,
$\alpha(0)=1$, $\alpha(\pi)=-1$, $\alpha\left(\frac{\pi}{2}\right)=0$.
The general function $\alpha(q)$ has to be determined from the
behaviour of the functions $f_A$, $f_B$ at $r\rightarrow\infty$.
We first consider the sphaleron and its boundary loop for $\lambda=1$
which happens to be a particularly simple case in the sense that the
energy along the loop can be calculated analytically. Choosing
\begin{equation}
f_B\equiv 0, \qquad H\equiv \frac{\xi}{2\eta^2}
\label{self1}
\end{equation}
simplifies the equations (\ref{leins}--\ref{lvier}) to
\begin{eqnarray}
\left(\frac{f_A^{\prime}}{r}\right)^{\prime} & = & \frac{2\eta^2}{r}K^2f_A
\label{reins} \\
K^{\prime} & = & -\frac{1}{r}Kf_A
\label{rzwei} \\
f_A^{\prime} & = & \eta^2 r \left(1-\frac{\xi^2}{4\eta^2}-K^2\right)
\label{rdrei} \\
(rK^{\prime})^{\prime} & = & \frac{f_A^2}{r}K-
\eta^2rK\left(1-\frac{\xi^2}{4\eta^2}-K^2\right)
\label{rvier}
\end{eqnarray}
while the boundary conditions (\ref{zero},\ref{infty})
for the remaining functions $f_A$, $K$ become
\begin{eqnarray}
K(0) = 0 \qquad & & \quad K(\rightarrow\infty)=
\pm\sqrt{1-\frac{\xi^2}{4\eta^4}} \nonumber \\
f_A(0) = -1, \quad & & \quad f_A(r\rightarrow\infty)=0
\label{selfbound}
\end{eqnarray}
as $\cos q=\frac{\xi}{2\eta^2}$ from the boundary condition on
$H\equiv \frac{\xi}{2\eta^2}$ at $r\rightarrow\infty$.
It is easy to check that eqs.\ (\ref{rzwei}) and (\ref{rdrei}) are first
integrals of eqs.\ (\ref{reins}) and (\ref{rvier}), hence a consistent solution
of the system (\ref{reins}--\ref{rvier}) can be found by solving the
first--order system (\ref{rzwei},\ref{rdrei}).
Solutions of eqs.\ (\ref{rzwei},\ref{rdrei}) saturate the inequalities
\begin{eqnarray}
\frac{1}{4r}\left[f_A^{\prime} -
\eta^2 r \left(1-\frac{\xi^2}{4\eta^2}-K^2\right)\right]^2 & \ge & 0
\nonumber \\
\frac{\eta^2}{2}r\left[K^{\prime} -\frac{1}{r}Kf_A\right]^2 & \ge & 0.
\label{self2}
\end{eqnarray}
The cross terms of the squares yield a lower bound for the Hamiltonian
(\ref{b5}) for $\lambda=1$ and $f_B\equiv0$, $H\equiv\frac{\xi}{2\eta^2}$:
\begin{eqnarray}
H_0\big|_{\lambda=1,f_B\equiv0,H\equiv\frac{\xi}{2\eta^2}} & = &
2\pi\left\{\frac{1}{4r}f_A^{\prime 2} +
\frac{\eta^2}{2}\left[rK^{\prime 2}+\frac{1}{r}(Kf_A)^2
\right]
+\frac{\eta^4}{4}r\left(1-\frac{\xi^2}{4\eta^4}-K^2\right)^2\right\}
\label{selfham} \\
& \ge & \pi\eta^2 \frac{d}{dr}
\left\{f_A\left(1-\frac{\xi^2}{4\eta^4}-K^2\right)\right\}
\label{self3}
\end{eqnarray}
The resulting Hamiltonian (\ref{selfham}) is exactly of the same form as
the one dimensional reduced Lagrangian (\ref{abhiggsvortex}) of the
Abelian Higgs model, which becomes obvious by replacing the functions
$(a,h)$ by $(f_A ,K)$ formally, and by replacing
$1$ by $1-\frac{\xi^2}{4\eta^4}$ in the potential. Solutions to
this system (which saturate the bounds (\ref{self2}) are known numerically.
Exploiting the boundary behaviour (\ref{selfbound}) of these solutions,
one can find their energy depending on the Lagrange multiplier
$\xi$ from the saturated lower bound (\ref{self3}),
\begin{equation}
{\cal E} = \int H_0\big|_{\lambda=1,f_B\equiv0,H\equiv\frac{\xi}{2\eta^2}} dr
=\pi\eta^2\left(1-\frac{\xi^2}{4\eta^4}\right).
\label{self4}
\end{equation}
Also the integral in the Chern--Simons number (\ref{so3ncs}) depends only
on the boundary values for this particular loop, and we obtain
\begin{equation}
{\cal N}_{CS}=\frac12\left(1-\frac{\xi}{2\eta^2}\right) \: ,
\label{self5}
\end{equation}
resulting in an analytic expression for the energy along the boundary loop,
\begin{equation}
{\cal E}({\cal N}_{CS}) = 4\pi\eta^2{\cal N}_{CS}\left(1-{\cal N}_{CS}\right)\: .
\label{self6}
\end{equation}
As a result the slopes of ${\cal E}({\cal N}_{CS})$ at the beginning and the end of the loop are
\begin{equation}
\left.\frac{\partial{\cal E}({\cal N}_{CS})}{\partial{\cal N}_{CS}}\right|_{{\cal N}_{CS}=0}=4\pi\eta^2,\qquad
\left.\frac{\partial{\cal E}({\cal N}_{CS})}{\partial{\cal N}_{CS}}\right|_{{\cal N}_{CS}=1}=-4\pi\eta^2.
\label{slopes}
\end{equation}
For coupling constants $\lambda>1$, the full set of equations
(\ref{leins}--\ref{lvier}) has to be be solved
numerically, using $\rho=\sqrt{2}\eta r$ as rescaled dimensionless radial
variable and eliminating $q$ from the boudary conditions at infinity.
In practice we replaced the boundary conditions (\ref{infty})
which also involve
the unknown function $\alpha(q)$ by
\begin{eqnarray}
H^2+K^2 & \stackrel{r\rightarrow\infty}{\longrightarrow} & 1 \nonumber \\
Kf_A+Hf_B & \stackrel{r\rightarrow\infty}{\longrightarrow} & 0
\nonumber \\
Hf_A^{\prime}+Kf_B^{\prime} & \stackrel{r\rightarrow\infty}{\longrightarrow} &
0 \nonumber \\
f_A^{\prime} & \stackrel{r\rightarrow\infty}{\longrightarrow} &
0
\label{varinfty}
\end{eqnarray}
and solved eqs.\ (\ref{leins}--\ref{lvier}) for given value of $\xi$,
identifying $q$ and $\alpha(q)$ from the asymptotic behaviour of the
numerical solutions afterwards.
The energy ${\cal E}$ and
Chern--Simons number ${\cal N}_{CS}$ can be computed for these solutions. This finally
yields a ``static minimal energy loop'' connecting two topologically
neighbouring vacua through the sphaleron configuration which corresponds
to the vortex of the Abelian--Higgs model. The energy along the loop
which can be parametrised by the Chern--Simons number, ${\cal E}({\cal N}_{CS})$, is
shown in Fig.\ \ref{loopfig1}, for several values of the ratio
$\frac{m_H^2}{m_W^2}=\lambda>1$. In this parameter region, the
Lagrange multipier is restricted to $|\xi|<\xi_c=2\eta^2$.
${\cal E}({\cal N}_{CS})$ approaches the vacua ${\cal E}=0$ for
$\xi\rightarrow\mp 2\eta^2$ with slopes
\begin{equation}
\left.\frac{\partial{\cal E}({\cal N}_{CS})}{\partial{\cal N}_{CS}}\right|_{{\cal N}_{CS}=0;1}=\pm4\pi\eta^2
\label{sslop}
\end{equation}
independently of $\lambda$, because the slope of the energy curve approaching
the vacua is given by $\pm 2\pi\xi_c$ \cite{AKY}.
\begin{figure}
\begin{center}
\includegraphics[bb=2cm 8.5cm 19cm 23cm,angle=0]{loopfig1.ps}
\end{center}
\caption{The energy along the boundary loop, ${\cal E}({\cal N}_{CS})$, for
$\lambda=1.0;1.2;1.6;2.0;3.0;4.0$.
The straight lines show the slopes of the energy curve at
the vacua ${\cal N}_{CS}=0,1$.}
\label{loopfig1}
\end{figure}
The symmetry ${\cal E}({\cal N}_{CS})={\cal E}(1-{\cal N}_{CS})$ is related to the invariance
of $H_0$ and $F_0$ in eqs.\ (\ref{b5}.\ref{withmult}) under the transformation
\begin{eqnarray}
H \mapsto -H \qquad & & \qquad f_A \mapsto \quad\! f_A \nonumber \\
K \mapsto\quad K \qquad & & \qquad f_B \mapsto -f_B
\label{symme}
\end{eqnarray}
which requires
\begin{equation}
q\mapsto \pi-q,\qquad \alpha(\pi-q)=\alpha(q).
\label{symme1}
\end{equation}
It is easy to check that this transfrmation yields
\begin{equation}
{\cal N}_{CS}\mapsto 1-{\cal N}_{CS}
\label{symme2}
\end{equation}
while the energy is invariant, yielding the symmetry of the energy along the
boundary loop mentioned above.
The boundary loop of the $SO(3)$--Higgs model we constructed
is different from the original AKY construction in the
Weinberg--Salam model \cite{AKY}. The Weinberg--Salam sphaleron has a pure
gauge connection
at infinity. This allowed a boundary loop construction \cite{AKY}
where all gauge fields
along the loop tend to a pure gauge, in particular, gauge and Higgs fields at
infinity could be constructed as gauge transformation of the initial
vacuum configuration, using the gauge transformation which rotates the
parameter fields and allows to gauge $f_C\equiv 0$. This gauge transformation
$g_{\Lambda}$ (\ref{gGauge}) also exists in the $SO(3)$--Higgs boundary loop
construction, but it can not be used to construct the loop configurations at
infinity from the initial vacuum. In fact, applying $g_{\Lambda=q}$ to
the initial vacuum $(\tilde{\Phi},\tilde{A}_i)_{q=0}=(\Phi,A_i)_+$ yields the
correct behaviour of the Higgs field, but not of the gauge field along
the loop at spatial infinity,
\begin{eqnarray}
(\tilde{\Phi},\tilde{A}_i)_{q=0} & = & \left(i\frac{\eta}{\sqrt{2}}\sigma_3,0\right)
\nonumber \\
& \stackrel{g_q}{\mapsto} &
\left(i\frac{\eta}{\sqrt{2}}(\cos q \sigma_3 + \sin q \hat{x}_i \sigma_i),
\frac{1-\cos q}{r}\epsilon_{ik}\hat{x}_k\left(-\frac{i}{2}\sigma_3\right)
-\frac{\sin q}{r}
\epsilon_{ik}\hat{x}_k\left(-\frac{i}{2}\hat{x}_i\sigma_i\right)\right) \ ,
\nonumber \\
\label{atinfinity}
\end{eqnarray}
as one can see by comparing with the Higgs and gauge fields along the loop at
spatial infinity, i.~e.
\begin{eqnarray}
(\tilde{\Phi},\tilde{A}_i)_q \quad & \vspace{-1cm} \stackrel{r\rightarrow\infty}{\sim} &
\nonumber \\
& & \hspace*{-1.5cm}
\left(i\frac{\eta}{\sqrt{2}}(\cos q \sigma_3 + \sin q \hat{x}_i \sigma_i),
\frac{1-\alpha(q)\cos q}{r}
\epsilon_{ik}\hat{x}_k\left(-\frac{i}{2}\sigma_3\right)
-\frac{\alpha(q) \sin q}{r}
\epsilon_{ik}\hat{x}_k\left(-\frac{i}{2}\hat{x}_i\sigma_i\right)\right) \ .
\nonumber \\
\label{gaugeloop}
\end{eqnarray}
Thi is due to the fact that the sphaleron, which has to be a loop
configuration, is one half times a pure gauge at infinity in accordance
with the
corresponding behaviour of the instanton.
We expect that this is a general feature of models which support
both instanton and
sphaleron, i.e.\ the sphaleron behaves like the instanton at spatial infinity,
a fact which must be reflected in the boundary loop construction.
\subsection{General vorticity sphalerons and geometrical loop construction}
\label{so3manton}
In general, the Abelian Higgs vortex has vorticity or winding number
$N\in {\rm I \hspace{-1mm} N}$. So far, we considered only $N=1$ for the $SO(3)$ sphaleron, but
the generalisation to sphalerons of arbitrary integer vorticity can
easily be achieved. We use this generalisation in our presentation of
the second, geometrical loop construction.
The starting point of the geometrical loop construction is the sphaleron
ansatz which we generalise to vorticity $N$, replacing the spatial unit
vector $\hat{x}$ by $\hat{n}=(\cos N\phi,\sin N\phi)$ in the ansatz,
\begin{equation}
\Phi=i\frac{\eta}{\sqrt{2}} h(r) \hat{n}_i\sigma_i,\qquad
A_i=\frac{N}{r}f(r)\epsilon_{ik}\hat{x}_k\left(-\frac{i}{2}\sigma_3\right),
\label{b6}
\end{equation}
where we use $h(r)$ and $f(r)$ as parameter functions \cite{Manton}.
Inserting the ansatz into the Hamiltonian (\ref{b2}) yields the radial
subsystem Hamiltonian
\begin{equation}
H_{sph}=
2\pi\left\{\frac14\frac{N^2}{r}f^{\prime 2}+\frac{\eta^2}{2}\left[
r h^{\prime 2}+\frac{N^2}{r}h^2(1-f)^2\right]+\frac{\lambda\eta^4}{4}
r(1-h^2)^2\right\}.
\label{b7}
\end{equation}
Finite energy and regularity at the origin require
\begin{equation}
f(r\rightarrow\infty)=1,\quad h(r\rightarrow\infty)=1,\qquad
f(0)=0,\quad h(0)=0.
\label{rand}
\end{equation}
and the solutions are equivalent to the vorticity $N$ Abelian Higgs vortices
as expected.
The geometrical loop construction starts from requiring finite energy, hence
the potential term in the Hamiltonian (\ref{b2}) forces
$|\Phi|=\eta\Rightarrow \Phi\in S^2_{Higgs}$ at spatial infinity
$({\rm I \hspace{-0.9mm} R}^2)^{\infty}=S^1_{space}$, described by the angular coordinate $\phi$.
Therefore we consider the Higgs field for $r\rightarrow\infty $
and add a single loop parameter $\tau\in S^1_{loop}$ to construct a nontrivial
mapping,
\begin{equation}
\Phi^{\infty}:S^1_{space}\times
S^1_{loop}\rightarrow S^2_{Higgs}.
\label{b7a}
\end{equation}
The simplest choice for this mapping with vorticity $N$ is
\begin{equation}
\Phi^{\infty}=i\frac{\eta}{\sqrt{2}}\vec{p}\cdot\vec{\sigma}, \qquad
\vec{p}=\left(\begin{array}{c}
\sin\tau\cos N\phi\\ \sin^2\tau\sin N\phi + \cos^2 \tau \\
\sin\tau\cos\tau(\sin N\phi -1)
\end{array}\right)
\label{b8}
\end{equation}
This fixes also the gauge field at infinity since the covariant derivative
in the Hamiltonian (\ref{b2}) has to vanish, requiring
\begin{equation}
A_i^{\infty}:=A_i(r\rightarrow\infty)=-\frac{1}{2\eta^2}[\Phi^{\infty},
\partial_i\Phi^{\infty}].
\label{b8a}
\end{equation}
According to the results of Section \ref{so3aky}, the gauge field at infinity
along the loop is not a pure gauge, in particular, it tends to half a pure
gauge for $r\rightarrow\infty$ for the sphaleron, $\tau=\frac{\pi}{2}$.
For $\tau=0$ and $\tau=\pi$, the loop has to start and end in the
vacuum which for convenience \cite{Manton} we choose to be
\begin{equation}
(\Phi,A_i)_{vac}=\left(i\frac{\eta}{\sqrt{2}}\sigma_2,0\right)
\label{b8b}
\end{equation}
in contrast to the choice of the vacua in Section \ref{so3aky}. This yields
the geometrical NCL ansatz
\begin{equation}
(\bar{\Phi},\bar{A}_i)_{\tau} = \left(i\frac{\eta}{\sqrt{2}}[(1-h(r))\vec{t}+h(r)
\vec{p}],-\frac{1}{2\eta^2}f(r)A_i^{\infty}\right)
\qquad {\rm with} \ \
\vec{t}=\left(\begin{array}{c}
0\\ \cos^2\tau\\
-\sin\tau\cos\tau
\end{array}\right)
\label{b9}
\end{equation}
The boundary conditions (\ref{rand}) then ensure
\begin{equation}
(\bar{\Phi},\bar{A}_i)_{\tau}\stackrel{r\rightarrow\infty}{\longrightarrow}
(\Phi^{\infty},A_i^{\infty}) \ ,
\label{bbb9}
\end{equation}
and $\vec{t}$ is chosen such that
\begin{equation}
(\bar{\Phi},\bar{A}_i)_{\tau=0}=(\bar{\Phi},\bar{A}_i)_{\tau=\pi}=(\bar{\Phi},\bar{A}_i)_{vac} \ .
\label{equation}
\end{equation}
One can also
check that $\Phi^2$ is a radial function of $r$ only which is necessary
for the consistency of the ansatz.
Inserting the loop ansatz (\ref{b9})
into the static Hamiltonian (\ref{b2}) yields
\begin{equation}
H_{\tau}=
2\pi\sin^3\tau\left\{\frac14\frac{N^2}{r}f^{\prime 2}+\frac{\eta^2}{2}\left[
r h^{\prime 2}+
\frac{N^2}{r}\left(h-f(h\sin^2\tau+\cos^2\tau)\right)^2
\right]+\frac{\lambda\eta^4}{4}\sin^2\tau r(1-h^2)^2\right\}.
\label{b10}
\end{equation}
For $\tau=\frac{\pi}{2}$, this geometrical
loop ansatz reduces to the sphaleron
ansatz (\ref{b6}), and of course $H_{\tau=\frac{\pi}{2}}=H_{sph}$, whereas
$H_{\tau=0}=0=H_{\tau=\pi}$ for the vacua at the beginning and the end of the
loop.
$H_{\tau}$ can be minimized numerically
for any value $\tau\in[0,\pi]$, yielding again a
``static minimal energy path'' connecting two neighbouring vacua through the
sphaleron, the energy along the loop being ${\cal E}(\tau)=\int H_{\tau}dr$.
One can also calculate the Chern--Simons number along the NCL, treating
$\tau=\tau(t)$ as time dependent with $\tau(-\infty)=0$, $\tau(\infty)=\pi$.
For the Manton loop ansatz (\ref{b9}), both the volume and the surface
integrals in eq.\ (\ref{bCS}) contribute, resulting in
\begin{equation}
{\cal N}_{CS}(\tau) =\frac{N}{2}\cos\tau
\sin^2\tau\int f'(1-h) dr + \frac{N}{2}(1-\cos\tau).
\label{b10a}
\end{equation}
For the sphaleron $\tau=\frac{\pi}{2}$, only the surface integral contributes
to ${\cal N}_{CS}=\frac{N}{2}$, whereas the loop as a whole has ${\cal N}_{CS}=N$.
The vorticity $N$ sphaleron therefore is the saddle point of a loop
connecting to vacua with Chern--Simons number difference $N$.
The ``static minimal energy path'' of the $SO(3)$ gauge Higgs
sphaleron in the geometrical construction, ${\cal E}({\cal N}_{CS})$, is shown in Fig.\
(\ref{loopfig2}) with vorticity $N=1$ and several values of $\lambda$.
Comparing the $\lambda=1$ geometrical loop with the corresponding boundary
loop, we find that the energy along the geometrical loop is lower than that
along the boundary loop except at the sphaleron and the vacua where they are
equal. This result holds also for $\lambda>1$.
Fig.\ \ref{loopfig3} shows the geometrical loop for several values of the
vorticity.
\begin{figure}
\begin{center}
\includegraphics[bb=2cm 8.5cm 19cm 23cm,angle=0]{loopfig2.ps}
\end{center}
\caption{The energy along the geoemtrical loop, ${\cal E}({\cal N}_{CS})$, with vorticity
$N=1$ and $\lambda=1.0;1.2;1.6;2.0;3.0;4.0$, and the energy along the
boundary loop for $N=1$, $\lambda=1.0$}
\label{loopfig2}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[bb=2cm 8.5cm 19cm 23cm,angle=0]{loopfig3.ps}
\end{center}
\caption{The energy along the geometrical loop,
$\frac{{\cal E}\left(\frac{{\cal N}_{CS}}{N}\right)}{N}$, with
$\lambda=1.0$ and vorticity $N=1;2;3;4;5$}
\label{loopfig3}
\end{figure}
\subsection{Magnetic and electric properties of Georgi--Glashow sphalerons}
Since the effective equations governing the sphaleron solution of the 3
dimensional Georgi--Glashow model are those of the Abelian Higgs vortex with
vorticity $N$ (hence also magnetic charge), we might
expect that the sphaleron itself may carry magnetic charge.
The electromagnetic field strength of the 3 dimensional Georgi--Glashow model
is the well-known 't Hooft electromagnetic tensor~\cite{tHP,Copeland}
\begin{equation}
{\cal F}_{\mu\nu}:=-\frac{\mbox{tr}[\Phi F_{\mu\nu}]}{\sqrt{-\mbox{tr}[\Phi^2]}} =
\frac{\vec{\phi}\cdot\vec{F}_{\mu\nu}}{|\vec{\phi}|} \: .
\label{b11}
\end{equation}
Inserting the sphaleron configuration (\ref{b6}) into (\ref{b11}) yields
${\cal F}_{\mu\nu}\equiv 0$, therefore the sphaleron itself carries neither
magnetic nor electric charge. The situation is different if we consider
the complete NCL as the configuration dependent on the time parameter.
For example, along the boundary loop presented in Section \ref{so3aky} with
$q=q(t)$ we have nonvanishing electric and magnetic fields,
\begin{eqnarray}
B & = & {\cal F}_{12} = -\frac{1}{r}\frac{Hf_A^{\prime}+Kf_B^{\prime}}
{\sqrt{H^2+K^2}} \label{fmag} \\
E_i & = & {\cal F}_{0i} = -\epsilon_{ij}
\frac{\hat{x}_j}{r}\frac{H\dot{f}_A+K\dot{f}_B}
{\sqrt{H^2+K^2}}. \label{fel}
\end{eqnarray}
It follows, by symmetry, that the electric charge of the NCL,
\begin{equation}
Q_e := \lim_{r\rightarrow\infty} \int_{S^(r)} \vec{E}\cdot d\vec{S},
\label{b15}
\end{equation}
is always zero, whereas the NCL configurations may acquire magnetic
charge
\begin{equation}
Q_m := \int B d^2x
\label{b15aa}
\end{equation}
at some values of the time parameter. Clearly, the magnetic charge of the
sphaleron and the initial and final vacua vanish according to the
values that the functions $f_A ,f_B , K$ and $H$ take for these configurations.
The same conclusion is arrived at also by inspecting the magnetic flux of the
vorticity $N$ configuration calculated in terms of the geometric loop
parameter $\tau$
\begin{equation}
\label{b15aaa}
Q_m = \sin^2 \tau \: \cos \tau \: N\: \int_0^{\infty}(1-h)f' dr \: ,
\end{equation}
which vanishes at the sphaleron, $\tau =\frac{\pi}{2}$ and the vacuua
$\tau =0$ and $\tau =\pi$.
One might think at this point that there is no reason why there should be no
nonvanishing electric flux if the temporal gauge condition $A_0 =0$ were
relaxed in the spirit of the Julia-Zee dyon~\cite{JuliaZee}. In the
$A_0 \neq 0$ one can solve the full Euler-Lagrange equations in Minkowski
space, of
\begin{equation}
{\cal L}_M=\mbox{tr}\left[
-\frac14 F_{\mu\nu}F^{\mu\nu} + \frac12 D_{\mu}\Phi D^{\mu}\Phi -
\lambda\left(\phi^2+\frac{\eta^2}{2}\right)^2\right],
\label{b12}
\end{equation}
in the static limit using the radially symmetric restriction of the fields
according to the Ansatz
\begin{eqnarray}
\Phi & = & i \frac{\eta}{\sqrt{2}} h(r) \hat{x}_i\sigma_i \nonumber \\
A_0 & = & i \frac{\eta}{\sqrt{2}} g(r) \hat{x}_i\sigma_i,
\qquad A_i= \frac{f(r)}{r} \epsilon_{ij}\hat{x}_j
\left(-\frac{i}{2}\sigma_3\right).
\label{b13}
\end{eqnarray}
To find solutions that lead to nonvanishing electric flux $Q_e$,
the solution must have the following asymptotic behaviour
\begin{equation}
g(r) \sim d_g + \sqrt{2}\pi\eta Q_e \ln r, \qquad m_W r \gg 1 \: ,
\label{b14}
\end{equation}
where $d_g$ is a constant. The asymptotic analysis of the equations, not
exibited here, indeed confirms the behaviour (\ref{b14}). These equations have been
integrated in \cite{Copeland} numerically.
However, this static electrically charged solution can not be called a
sphaleron since the logarithmic behaviour of $g(r)$
(which is a consequence of $\ln r$ being the fundamental solution of the
Laplace operator in two
dimensions) destroys the integrability of the static Hamiltonian, i.e.\
this electrically charged classical configuration does not have finite
energy, rendering it useless as a pseudoparticle in
quantum field theory. Technically, this is related to the well--known fact
that the energy integral of point charges in two dimensional electrodynamics
is divergent.
\section{Summary and discussion}
We have analysed the first two in the hierarchy of
$SO(d)$ gauged $d$ dimensional Higgs models
where the Higgs fields are in the $d$-dimensional
vector representation of $SO(d)$. In $d=2$ and
$3$ these are the familiar Abelian Higgs model and
the Georgi-Glashow model respectively. The
reason for having chosen these models is that they
both support topologically stable finite
action solutions which we interpret as the instantons,
and, they are expected to support
sphaleron solutions in the static limit, which we find
indeed to be the case. Thus we have
analysed the first two Higgs models in this hierarchy,
which support both instantons and
sphalerons. This is the main criterion of the work, and
our motivations for it are explained in
the Introduction. Our analysis points the way to tackling
the $d=4$ case~\cite{OT} which is of
some definite physical relevance inasfar as it promises a
Coulomb instanton gas~\cite{so4inst},
but which is considerably more complex.
Since the instantons of these two models are well known
solutions, namely the Nielsen--Oleson
vortices~\cite{NO} and the 't Hooft--Polyakov
monopole~\cite{tHP}, the analysis of this paper
is restriced to the study of the sphalerons of these models.
Indeed, the sphalerons of the
Abelian Higgs model have been studied extensively in the
literature~\cite{BS,C,FH}, but we give
our own version here. The reason for this is firstly so
that the construction of the sphalerons
of both models should proceed in the same lines,
especially since one of criteria is to map the
way to carry this analysis to the $d=4$ case.
Secondly, our construction of the sphaleron of
the Abelian Higgs model follows exactly the same
procedure as those used in the corresponding
work for the Weinberg-Salam model, namely that of
constructing a NCL carried out by Manton
\cite{Manton}, and also that of constructing the
path of finite energy configurations carried
out by Akiba {\it et.al.}~\cite{AKY}. This contrasts with
the presentation in the literature~\cite{BS,C,FH}
where the procedure is quite different from the
case of the Weinberg-Salam model~\cite{Manton,
Klinkhamer, AKY}, in particular imposing periodic
boundary conditions~\cite{BS} unnecessarily.
This part of our work is presented in Section 2,
where we have given both construction of the
sphaleron, namely that of Manton and
Klinkhamer~\cite{Manton, Klinkhamer} and that of Akiba
{\it et.al.}~\cite{AKY}. In addition, we have made a contrast
of our procedures with those existing in
the literature. Our procedure here making it possible
to formulate this problem in complete
parallel to the Weinberg-Salam case is made possible by
our use of the systematically
derived~\cite{desc} Chern-Pontryagin density used in
calculating the Chern-Simons number.
In the larger part of this work we have presented the
construction of the sphaleron in the
three dimensional Georgi-Glashow model. This is the work
in Section 3. Again we have presented
the construction of the sphaleron in both the Manton and
Klinkhamer~\cite{Manton, Klinkhamer}
procedure and the Akiba {\it et.al.}~\cite{AKY} one. It is very
interesting that in the former case
\cite{Manton, Klinkhamer}, we find that the sphaleron is
a solution to the effective Abelian
Higgs model in the two spatial dimensions, embedded in the
$SO(3)$ model. In that case we
have constructed a family of sphalerons characterised by a
vortex number $N$, availing of the
fact that radially symmetric fields in two dimensions have
integer vorticity. We have also
inquired whether this sphaleron has magnetic flux related
to its vorticity $N$ and have found
that the sphaleron itself, as well as the vacuua it falls
between, have zero magnetic flux,
but that intermediate field configurations do have
nonvanishing magnetic flux. We have also
verified that by relaxing the temporal gauge, one can find
solutions to the static field
equations leading to a nonvanishing electric
flux~\cite{Copeland}, but in that case the energy
diverges logarithmically and hence there is no sphaleron
solution that can be gainfully employed
in semiclassical field theory. In addition to this, we
have pursued the construction in the
procedure of Akiba {\it et.al.}~\cite{AKY} and have constructed
the finite energy path interpolating
between the two vacuua with the sphaleron at the top of
this path with Chern--Simons number equal
to $1\over 2$. An interesting circumstance here is that
when the coupling constant of the Higgs
self-interaction potential in the Georgi-Glashow model
takes the critical value for which the
(static) embedded Abelian Higgs model can saturate the
corresponding Bogomol'nyi bound, the
finite energy path can be constructed analytically without
recourse to numerical computations.
Our analysis of these two models has clarified the
similarities and differences between the
sphalerons of the Weinberg-Salam model and those of the
models in our hierarchy of $SO(d)$ Higgs
models. In this respect, the $d=3$ case is the most enlightening. We have
learnt that the structure of the geometric loop~\cite{Manton,Klinkhamer}
behaves very much like that of the Weinberg-Salam model, and that the
boundary loop~\cite{AKY} also behaves similarly inspite of the fact that the
actual boundary conditions are quite different in the two cases. In the process,
we have given a unified treatment for the construction of energy loops
interpolating between vacuua for all models, bringing the treatment of the Abelian
Higgs case into line with the others. This we have done employing both types of
energy loops, geometrical~\cite{Manton} and boundary~\cite{AKY}, whence we have
learnt that the geometric loop is lower than the boundary loop everywhere except
at the sphaleron and the vacuua where they are equal. Presumably this is the case
also for the Weinberg-Salam model.
As a final remark we emphasise the similarity between the boundary loop
construction for the Winberg-Salam and the Georgi-Glashow models --
in both of them the functions $f_B$ and $H$ are excited on the boundary loop.
In addition both theories tend to sigma models in the limit of the Higgs masses
becoming infinit. With these two similarities in place, it seems that there may
well be bisphalerons \cite{BK,J} in the static 3 dimensional
Georgi-Glashow model.
\bigskip
\noindent
{\bf Acknowledgments}
\noindent
This work was carried out in part under
Basic Science Research project SC/97/636 of
FORBAIRT. FZ acknowledges a Presidency of Ireland Postdoctoral
Fellowship from the HEA. We are very grateful to Yves Brihaye for
valuable discussions.
|
\section{Introduction}
Consider a quantum field theory with an n-dimensional vacuum manifold
${\cal M}_n$. An illustrative example which will be made use of in
most of this Letter is a theory of four real scalar fields $\phi_i,
\,\, i = 1, .., 4$, with a zero temperature potential
\begin{equation} \label{toy}
V(\phi) \, = \, {1 \over 4} \lambda \bigl( \sum_{i=1}^4 \phi_i^2 -
\eta^2 \bigr)^2
\end{equation}
which has as its vacuum manifold the three-dimensional sphere
${\cal M}_3 \, = \, S^3 \, = \, \{ (\phi_i): \sum_{i=1}^4 \phi_i^2 =
\eta^2 \}$.
By freezing out certain combinations of the original fields we obtain
a sub-manifold ${\cal M}_m, \, m < n,$ of the original vacuum manifold
${\cal M}_n$. In the above example we can set $\phi_3 = \phi_4 = 0$
and obtain the one-dimensional sub-manifold
${\cal M}_1 \, = \, S^1 \, = \, \{ (\phi_i): \sum_{i=1}^2 \phi_i^2 =
\eta^2 \}$
As is well known$^{\cite{Kibble}}$, the topology of the vacuum
manifold determines the types of topological defects which the theory
admits. If the topology of ${\cal M}_m$ is such that k-dimensional
defects are possible (in four-dimensional space-time the criterion is
$\Pi_{2 - k}({\cal M}_m) \, \neq \, {\bf 1}$,
where $\Pi_l$ is the l'th homotopy group), then it is possible to
construct configurations of the unconstrained fields which correspond
to k-dimensional defects. If these configurations satisfy the
equations of motion of the unconstrained theory, we have an embedded
defect$^{\cite{TB,T92,BTB}}$. In the above example, the theory of
$(\phi_1, \phi_2)$ and vacuum manifold ${\cal M}_1$ admits linear
defects called cosmic strings. These string solutions with $\phi_3 =
\phi_4 = 0$ satisfy the full set of equations of motion and are hence
embedded strings.
Embedded defects are not topologically stable. In general, they are
not dynamically stable, either$^{\cite{instab}}$. The defect
configurations can unwind by exciting the frozen field
combinations. In the above example, the embedded string can unwind by
escaping into the $\phi_3$ and $\phi_4$ field dimensions.
If they were stable during a certain period in the early Universe,
embedded defects could play an important role in cosmology. The first
important example arises in the Standard Model of strong interactions,
which below the confinement scale is described by a sigma model
analogous to our toy model of Eq. (\ref{toy}), with $\phi_1$ denoting
the $\sigma$ field and $\phi_i, i = 2,3,4$ the three pions. This
theory admits no stable defects. Since $\Pi_3({\cal M}_3) \neq {\bf 1}$
there are $\Pi_3$ textures$^{\cite{Davis}}$ (these are $k = -1$
dimensional defects in the above notation), but in the
absence of a stabilizing Skyrme term $\Pi_3$ textures are
unstable$^{\cite{Turok}}$. However, by setting the charged pion fields
$\phi_3 = \phi_4 = 0$, we can construct embedded strings (pion
strings$^{\cite{pion}}$). A pion string along the z axis is given by
the configuration
\begin{equation} \label{string}
\phi(r, \theta) \, = \, \eta \rho(r) e^{i \theta}
\end{equation}
where $r, \theta$ are the polar coordinates in the x-y plane and $\phi
= \phi_1 + i \phi_2$. The profile function $\rho(r)$ satisfies the
boundary conditions $\rho(r) \rightarrow 0$ for $r \rightarrow 0$ and
$\rho(r) \rightarrow 1$ for $r \rightarrow \infty$. In Ref. \cite{BZ}
it was shown that due to the anomalous coupling with electromagnetism,
pion strings may provide a mechanism for generating primordial
galactic magnetic fields.
A second application of stabilized embedded defects is to
baryogenesis. Since defects trap energy in the unbroken state and
since they are out-of-equilibrium field configurations, they can play
an important role in baryogenesis$^{\cite{BDH}}$. Defect-mediated
baryogenesis may be implemented at the GUT scale$^{\cite{BDH}}$, at
the electroweak scale$^{\cite{BD,BDPT}}$, and possibly even at the QCD
scale$^{\cite{BHZ}}$. Embedded defects can contribute to baryogenesis
in the same way$^{\cite{BD}}$ as topological defects. Consider, for
example, the standard electroweak theory based on the gauge group
$SU(2) \times U(1)$ with a complex Higgs doublet $\Phi = (\phi_+,
\phi_0)$, where the subscript indicates the electric charge. The
electroweak Z string$^{\cite{T92,Nambu}}$ is a solution of the
equations of motion obtained by embedding the
Nielsen-Olesen$^{\cite{NO}}$ $U(1)$ string in the following way:
\begin{equation}
\Phi \, = \, \rho_{NO}(r) e^{i \theta} (0, 1) \, , \,\,\,\, Z_{\mu} \, = \,
A_{\mu, NO} \, ,
\end{equation}
where $\rho_{NO}(r)$ and $A_{\mu, NO}$ are the scalar profile function
and the gauge field, respectively, of the Nielsen-Olesen string, and
$Z_{\mu}$ is the gauge field associated with the (electrically
uncharged) Z boson. It can be shown$^{\cite{instab}}$ that for
realistic values of the weak mixing angle, the electroweak string is
unstable at zero temperature. If stable, electroweak strings could
mediate electroweak scale baryon number violating
processes$^{\cite{BD}}$.
A third application of embedded defects is their role in mediating
interactions between stable topological defects. In Ref. \cite{ABES}
it has been shown that in a $O(3)$ linear sigma model there exists an
attractive force between monopoles and embedded walls and that upon
contact the monopole charge spreads out on the wall. Thus, embedded
walls can mediate a long range force between monopoles (they sweep up
monopoles and antimonopoles) and thus alleviate the monopole problem,
as proposed in the context of topological walls in
Ref. \cite{DLV}. The collapse of stable embedded defects could also
possibly lead to signatures in the X-ray and cosmic ray backgrounds.
There are many possible embedded defects - embedded
monopoles$^{\cite{BN}}$, embedded strings like the abovementioned pion
and electroweak strings, and embedded walls$^{\cite{BT}}$. In general,
the criteria for the existence of embedded defects are more
complicated$^{\cite{BTB}}$ than in our simple toy model. However, for
the rest of this Letter we shall stick to the simple model.
The formation of non-topological defects was studied in \cite{NY} and
\cite{ABD}. As shown in \cite{NY}, the formation probability of
electroweak strings is negligibly small if finite temperature effects
are neglected. In contrast, if embedded defects were stable at
temperatures below the symmetry breaking phase transition, they would
be produced in comparable abundance to that of topological defects of
the same dimension, by the usual Kibble argument$^{\cite{Kibble}}$.
As has been pointed out, it is possible that embedded defects can be
stabilized by bound states$^{\cite{VW,HHVW}}$ or by external electric$^{\cite{Periv}}$ and magnetic
fields$^{\cite{GM}}$. These mechanisms, however, are not generic. In
this Letter we propose a new and rather generic stabilization
mechanism which works whenever the fields excited in the embedded
defects are uncharged and the non-excited fields are charged, and when
the system is in a finite temperature plasma made up of charged
fields. In our examples, the charge will be the usual electric
charge. The mechanism, however, is more general and applicable to any
kind of charge, provided that the plasma is made up of particles which
carry that charge.
The basic idea is as follows: interactions with the charged plasma
will generate corrections to the effective potential for the scalar
fields which lift the potential in the directions of the charged
fields. This reduces the vacuum manifold ${\cal M}_n$ of the zero
temperature theory to a lower dimensional sub-manifold ${\cal M}_m$,
thus providing a way to stabilize embedded defects of the full theory
which are topological defects from the point of view of ${\cal M}_m$.
Our effect is different from an ordinary finite temperature effect. We
assume that the scalar fields are not in thermal equilibrium with the
plasma, but that the gauge fields which carry the charge force are, in
contrast to the usual framework of finite temperature field theory
where it is assumed that all fields are in thermal equilibrium. It
has been shown$^{\cite{HHVW}}$ that ordinary finite temperature
effects (which lift the effective potential in all field directions)
cannot in general stabilize embedded defects.
\section{Analytical Considerations}
As a toy model for the analytical study of the stabilization of
embedded defects by plasma effects we consider the chiral limit of the
QCD linear sigma model, involving the sigma field $\sigma$ and the
three pions ${\vec \pi} = (\pi^0, \pi^1, \pi^2)$, given by the
Lagrangian
\begin{equation} \label{lag1}
{\cal L}_0 \, = \, {1 \over 2} \partial_{\mu} \sigma \partial^{\mu}
\sigma + {1 \over 2} \partial_{\mu} {\vec \pi} \partial^{\mu} {\vec
\pi} - {\lambda \over 4} (\sigma^2 + {\vec \pi}^2 - v^2)^2 \, ,
\end{equation}
where $v^2$ is the ground state expectation value of $\sigma^2 + {\vec
\pi}^2$. In the following, we will denote the potential in
(\ref{lag1}) by $V_0$.
Two of the scalar fields, the $\sigma$ and $\pi_0$, are electrically
neutral, the other two are charged. Introducing the coupling to
electromagnetism, it is convenient to write the scalar field sector
${\cal L}$ of the resulting Lagrangian in terms of the complex scalar
fields
\begin{equation}
\pi^+ \, = {1 \over {\sqrt{2}}} (\pi^1 + i \pi^2) \, , \,\,\,
\pi^- \, = {1 \over {\sqrt{2}}} (\pi^1 - i \pi^2) \, .
\end{equation}
According to the minimal coupling prescription we obtain
\begin{equation} \label{lag2}
{\cal L} \, = \, {1 \over 2} \partial_{\mu} \sigma \partial^{\mu}
\sigma + {1 \over 2} \partial_{\mu} \pi^0 \partial^{\mu} \pi^0 +
D_{\mu}^+ \pi^+ D^{\mu -} \pi^- - V_0 \, ,
\end{equation}
where
\begin{equation} \label{partials}
D_{\mu}^+ \, = \, \partial_{\mu} + e A_{\mu} \, , \,\,\,\, D_{\mu}^-
\, = \, \partial_{\mu} - e A_{\mu} \, .
\end{equation}
If the gauge fields are part of the finite temperature plasma, we can
insert (\ref{partials}) into (\ref{lag2}) and do a Hartree-like
approximation by substituting
\begin{equation}
<A_{\mu}> \, = \, 0 \, , \,\,\,\, <A_{\mu}A^{\mu}> = \kappa T^2 \,
\end{equation}
where $T$ is the temperature and $\kappa$ is a numerical constant of
order unity, to extract from (\ref{lag2}) an effective potential for
the scalar fields of the form
\begin{equation} \label{poteff}
V_{T, eff}(\sigma, {\vec \pi}) \, = \, V_0 + \frac{1}{2}e^2\kappa T^2
((\pi^1)^2 + (\pi^2)^2) \, .
\end{equation}
At zero temperature, the vacuum manifold is $S^3$, but at finite
temperature it reduces to $S^1$. Thus, it is not unreasonable to
suspect that at high $T$ the embedded string (\ref{string}) is
stabilized.
To check the stability of the pion string (\ref{string}), we consider
the following variational ansatz for an unstable mode which obeys the
cylindrical symmetry of (\ref{string}) but escapes into the charged
field directions:
\begin{eqnarray} \label{var}
\phi = \sigma + i \pi^0 \, &=& \, v \rho(r) e^{i \theta} \, \,\,
\rho(r) \, = \, (1 - e^{- \mu r}) \\
\pi^1(r) \, &=& \, v \chi(r) \, , \nonumber
\end{eqnarray}
where $\mu$ is the width of the string (note that since the temperature only affects the charged fields and the embedded defects is made up of neutral
field, the width will be independent of $T$) and $\chi \ll 1$. We now
calculate the mass per unit length $I$ of the configuration
(\ref{var}) and compare it with the corresponding value $I_0$ for the
embedded string ($\chi = 0$). We obtain
\begin{equation}
I - I_0 \, = \, \pi v^2 \int_0^R dr r \chi^2 [e^2
\kappa T^2 - 2 \lambda v^2 e^{-\mu r}(1 - {1 \over 2} e^{-\mu r})] \, .
\end{equation}
The condition for stability is $I - I_0 > 0$ for all $\chi(r)$. As a
sufficient condition we get
\begin{equation} \label{bound}
T \, > \, 2 \lambda^{1/2} \kappa^{-1/2} e^{-1} v = T_D \, .
\end{equation}
If this temperature $T_D$ is smaller than the temperature $T_c$ of the
phase transition, then there will be a period $T \in [T_D, T_c]$
during which embedded defects are stable. The critical temperature
$T_c$ is obtained from the finite temperature effective
potential$^{\cite{FT}}$. Up to factors of order unity we obtain in our
toy model $T_c \simeq v$. Hence, provided that $\lambda e^{-2} \ll 1$
there is a period in the early Universe during which embedded defects
are stable.
So far, however, we have only studied the stability towards
perturbations maintaining the cylindrical symmetry. It is also
important to analyze the stability towards fragmentation. To do this,
we can consider the variational ansatz (\ref{var}) with a z-dependence
of $\rho$ and $\chi$. The only way this changes the previous analysis
is by adding to $I$ extra positive definite terms coming from spatial
gradient energies from z-derivatives. Hence, it is more difficult to
create local z-dependent perturbations than z-independent ones. Thus,
an embedded string with $T < T_D$ is also stable towards
fragmentation.
\section{Numerical Results}
We have simulated the formation of embedded defects in the presence of
a charged plasma using a code based on the one employed in
\cite{NSY}. Four scalar fields $\phi$ are evolved numerically on a
three-dimensional lattice by means of the equations of motion derived
from (\ref{lag1}) with the potential $V_0$ replaced by the effective
potential (\ref{poteff}). All dimensional quantities were rescaled by
appropriate powers of $v$ to make them dimensionless. For most
simulations, a box size of $50^3$ was used (we checked that the basic
results were insensitive to the box size by executing $100^3$ and
$200^3$ box simulations.). The spatial resolution was
$\Delta x = v^{-1}$, and the time steps were chosen as $\Delta t = {1
\over 100} \Delta x$. We used two different sets of initial
conditions. The results shown below are for {\it true vacuum} initial
conditions, choosing the fields to be randomly distributed (on a
length scale of $\Delta x$) over the entire vacuum manifold ${\cal
M}_n = S^3$:
$\vert \phi(x) \vert \, = \, v \, , \,\,\, {\dot \phi}(x) \, = \, 0$.
We also performed simulations with {\it false vacuum} initial
conditions determined by
$\phi(x) \, = \, 0 \, , \,\,\, \vert {\dot \phi}(x) \vert \, = \, T^2 \, ,$
with random orientations of the direction of ${\dot \phi}$ in the
field tangent space.
\begin{figure}
\begin{center}
\mbox{\psfig{figure=nagfig1.ps,height=8cm}}
\end{center}
\caption{Time evolution of the number $N$ of points where $\vert\phi\vert
<0.5v$. The simulations start at $t=1$ and the result
is the averaged value over 20 simulations for different initial
phase configuration in a $50^3$ box. The solid line corresponds to
$T=10^{-3}v$, the dotted one to $T=10^{-2}v$, the long-dashed
one to $T=10^{-1}v$, and the short-dashed one to $T=v$.}
\end{figure}
The presence of embedded defects implies the existence of points in
space with $\vert \phi(x) \vert < \alpha v$, where $\alpha < 1$ is an
arbitrary constant which determines what is meant by the core radius
of the defect (and chosen to be $\alpha = 1/2$ in Figs. 1 and 2). For
true vacuum initial conditions, we expect that the number $N$ of grid
points which satisfy the above criterion is initially independent of
the temperature $T$ (and simply indicative of the random fluctuations
set up by the initial gradient energy). As a function of time, we
expect the number of points to decrease if defects are unstable, but
to converge to a constant value if defects are stable. Thus, we expect
an abrupt increase in $N$ at later times at the stabilization
temperature $T_D$. We observed this in our simulations. Fig. 1 shows
the results for the case $\kappa = e = 1$ and $\lambda = 10^{-2}$. The
numerically determined value of $T_D$ lies between $10^{-2}v$ and
$10^{-1}v$, in good agreement with the analytical upper bound of
(\ref{bound}). The fact that there are points with $\vert \phi(x) \vert
< \alpha v$ for all temperatures is due to the large energy in the
system due to the initial gradient energy which leads to
fluctuations. In an expanding Universe, these fluctuations would be
damped. However, expansion is not included in our code.
We can introduce a damping term in order to reduce the effects of fluctuations.
In Fig. 2 we compare the three-dimensional distribution of
points with $\vert \phi(x) \vert < \alpha v$ in a simulation
with $T = v$ (stable defects expected) and with $T = 10^{-3}v$
(no stable defects expected). In the first case, although the zero
point constraint is more stringent than the second one due to
the smaller $alpha$, strings of grid
points with $\vert \phi(x) \vert < \alpha v$ are apparent, whereas
such strings are absent in the second case.
\begin{figure}
\begin{center}
\mbox{\psfig{figure=nagfig2a.ps,height=5cm}}
\quad\quad \mbox{\psfig{figure=nagfig2b.ps,height=5cm}}
\end{center}
\caption{The three-dimensional distribution of cells where
$\vert \phi \vert < \alpha v$ at $t=80$ for simulations including
the damping term. The left graph is for $T=v$, $\alpha=0.1$
and the right one for $T=10^{-3}v$, $\alpha=0.15$.}
\end{figure}
\section{Discussion}
We have discussed the stabilization of embedded defects made up of
neutral fields in the presence of a charged plasma. We illustrated the
effect in the context of pion strings which can be stabilized by the
electromagnetic plasma. However, the same mechanism applies to other charges,
for example to the color-charged plasma above the QCD confinement scale which has the potential of stabilizing color-neutral embedded defects.
The stabilization of neutral embedded defects is due to finite temperature plasma effects which add (to leading order) quadratic corrections to the effective potential for the charged fields. No external charged fields (as in the stabilization mechanism of \cite{Periv} in which embedded defects which couple to the external charge are stabilized by an induced angular momentum) are required.
For our mechanism to operate it is important that the fields which make up the embedded defects are no longer in thermal equilibrium whereas the gauge fields which carry the charges are still thermally excited. This condition is easier to realize for global defects such as pion strings than for local defects such as electroweak strings in which the gauge fields themselves play an important role.
Important questions for future study are the formation probability and length distribution of stabilized embedded strings. Unlike topologically stable strings, stabilized embedded strings may have ends (in which case they are
expected to be short-lived). Answers to these questions will be crucial in order to be able to discuss the cosmological consequences of the scenario. Work on this topic is in progress.
\centerline{Acknowledgments}
MN acknowledges the kind hospitality of Brown University and University of
British Columbia (UBC) where most of this work was done. R.B. thanks W. Unruh for hospitality at UBC. We are grateful to T. Vachaspati for comments on a draft of the paper.
The work of R.B. is supported in part by the U.S. Department of Energy under
Contract DE-FG02-91ER40688, TASK A.
|
\section{Introduction}
In the last few years many efforts have been made to understand the origin
of the cosmic X-ray background (CXB), discovered more than 35 years ago by
Giacconi et al. (1962).
One of the competing hypotheses - the truly diffuse emission origin
(see e.g. Guilbert and Fabian, 1986) -
has been rejected by the small deviation
of the cosmic microwave background spectrum from a blackbody shape
(Mather et al., 1994). Therefore only the alternative interpretation - the
discrete sources origin - is left.
Indeed, ROSAT deep surveys, reaching a source density of $\sim 1000$
deg$^{-2}$ at a limiting flux
of $10^{-15}$ ergs cm$^{-2}$ s$^{-1}\ $ (Hasinger et al., 1998; McHardy et al., 1998),
have already resolved the
majority (70$-$80 \%) of the soft ($E<2$ keV) CXB into discrete sources.
Spectroscopic observations (Shanks et al., 1991; Boyle et al.,
1993, 1994; McHardy et al., 1998; Schmidt et al., 1998) of the sources
with fluxes greater than $\sim 5 \times 10^{-15}$ ergs cm$^{-2}$ s$^{-1}\ $ have shown that the
majority ($50-80$ \%) of these objects are broad line AGN at $<z>
\sim 1.5$. An important minority ($10-20$ \%) of ROSAT sources are
spectroscopically identified with X-ray luminous Narrow Emission Line
Galaxies (Griffiths et al., 1995, 1996; McHardy et al., 1998), whose
real physical nature (obscured AGN, starburst) is at the moment debated
in the literature (see e.g. Schmidt et al., 1998).
Since their average X-ray spectrum is harder than that of the broad line
AGNs (Almaini et al., 1996; Romero-Colmenero et al., 1996) and similar
to that of the remaining unresolved CXB, these objects could also be
substantial contributors to the CXB at higher energies.
About 10\% of the
ROSAT sources are identified with clusters of galaxies (see e.g. Rosati et
al.,1998). Thus, it is clear that the ROSAT satellite has been
successful in resolving almost all the soft CXB into discrete sources.
Furthermore,
optical observations of the faint ROSAT sources has lead to the
understanding of the
physical nature of the objects contributing to it.
On the contrary, at harder energies, closer to
where the bulk of the CXB resides, the origin of the CXB
is still matter of debate.
Before the ASCA and Beppo-SAX satellites, which carry the first
long lived imaging instruments in the 2-10 keV energy band, surveys in
this energy range were made using passively collimated X-ray detectors
that, because of their limited spatial resolution, allowed the
identification of the brightest X-ray sources only, which represent a
very small fraction ($< 5\%$) of the CXB (Piccinotti et al., 1982).
The so called ``spectral paradox"
further complicates the situation: none of the single classes of
X-ray emitters in the Piccinotti et al., (1982) sample is characterized
by an energy spectral distribution similar to that of the CXB. Due to
the lack of faint, large and complete samples of X-ray sources selected
in this energy range, the contribution of the different classes of
sources to the hard CXB was evaluated through population
synthesis models and different classes of X-ray sources were proposed
as the major contributors by a number of authors
(e.g. starburst
galaxies, absorbed AGN, reflection dominated AGN,
see e.g. Griffiths and Padovani, 1990; Madau, Ghisellini and Fabian, 1994;
Comastri et al., 1995; Zdziarski et al., 1993).
Recent results from ASCA and Beppo-SAX observations of individual objects
and/or medium-deep survey programs
(Bassani et al., 1999; Maiolino et al., 1998a; Turner et al., 1998; Boyle et al.,
1998a; 1998b; Akiyama et al., 1998)
seem to
favor the strongly absorbed AGN hypothesis but deeper investigations
are still needed to confirm this scenario.
At the {\it Osservatorio Astronomico di Brera}, a serendipitous search
for hard (2-10 keV band) X-ray sources using data from the GIS2
instrument onboard the ASCA satellite is in progress (Cagnoni, Della
Ceca and Maccacaro 1998, hereafter Paper I; Della Ceca et al. 1999, in
preparation) with the aim of extending to faint fluxes the census of
the X-ray sources shining in the hard X-ray sky. The strategy of the
survey, the images and sources selection criteria and the definition of
the sky coverage are discussed in Cagnoni, Della Ceca and Maccacaro
(1998).
In Paper I, a first sample of 60 serendipitous X-ray sources, detected
in 87 GIS2 images at high
galactic latitude ($|b| > 20^o$) covering $\sim 21$ square deg.
was presented.
This sample has allowed the authors to extend the description of the
number-counts relationship down to a flux limit of $\sim 6\times
10^{-14}$ ergs cm$^{-2}$ s$^{-1}\ $ (the faintest detectable flux),
resolving {\bf directly} about 27\% of the (2-10 keV) CXB.
Here we study the spectral properties of the 60 ASCA sources listed in
Paper I, combining GIS2 and GIS3 data. We have carried out both an
analysis of the ``stacked" spectra of the sources in order to
investigate the variation of the source's average spectral properties
as a function of the flux, and a ``hardness-ratio" (HR) analysis of the
single sources. This latter method, which is equivalent to the
``color-color" analysis largely used at optical wavelengths, is
particularly appropriate when dealing with sources detected at a low
signal-to-noise ratio (e.g. Maccacaro et al., 1988; Netzer, Turner and
George, 1994). We have defined two independent HR and we have compared
the position of the sources in the HR diagram with a grid of
theoretical spectral models which are found to describe the X-ray
properties of known classes of X-ray emitters.
The paper is organized as follows. In section 2 we present the sample
and we define the ``Faint'' and the ``Bright'' subsamples. In section 3 we
present the data, we discuss the data analysis and we define the two HR
used. In section 4 we report the results of the stacked spectra and
of the HR analysis and we compare them with those expected from simple
spectral models and with the CXB spectra. Summary and conclusions are
presented in section 5.
\vfill
\eject
\section{Definition of the ``Faint" and ``Bright" Subsamples}
The basic data on the 60 X-ray sources used in this paper are reported in
Table 2 of Paper I.
To investigate if the spectral properties of the sources depend on
their brightness we have defined two subsamples according to the
``corrected'' 2-10 keV count rate (hereafter CCR)
\footnote{Due to the vignetting of the XRT and the PSF, a source with a
given flux will yield an observed count rate depending on the position
of the source in the field of view. With ``corrected count rate" we
mean the count rate that the source would have had if observed at some
reference position in the field of view and whitin a given extraction region.
The definition of the source extraction region and of the reference position in the
field of view, used to determine the corrected CCR, are discussed in section 3.2.
The CCR used here can be obtained dividing the
unabsorbed 2-10 keV flux reported in Paper I by the count rate to
flux conversion factor of 1cnt/s (2-10) = $11.46 \times 10^{-11}$ ergs cm$^{-2}$ s$^{-1}\ $.
This value is appropriate for a power law
model with energy spectral index of 0.7, filtered by a Galactic
absorbing column density of $3 \times 10^{20}$ cm$^{-2}$.}
.
The 20 brightest sources (CCR $\geq 3.9 \times 10^{-3}$ cts s$^{-1}\ $ ) define the
``Bright" subsample, while the remaining 40 sources define the ``Faint"
subsample.
The dividing line of $3.9 \times 10^{-3}$ cts s$^{-1}\ $ corrisponds to an
unabsorbed 2 - 10 keV flux of
$\sim 5.4\times 10^{-13}$ or $\sim 3.1\times 10^{-13}$ ergs cm$^{-2}$ s$^{-1}\ $
for a source described by a power law model with energy spectral
index of 0.0 and 2.0, respectively absorbed by a
Galactic column density of $3 \times 10^{20}$ cm$^{-2}$.
We note that the numbers reported above are a very weak function of the
Galactic absorbing column density along the line of sight (which ranges
from $10^{20}$ cm$^{-2}$ to $9 \times 10^{20}$ cm$^{-2}$ for the
present sample).
We prefer to use the CCR instead of the flux because the CCR is (once
the corrections due to the vignetting and the PSF have been applied) an
{\it observed} quantity and it is independent on the spectral
properties of the source; on the contrary the flux is model
dependent. In first approximation fainter CCR corresponds to fainter
sources.
\section{Data analysis}
All the ASCA images used in Paper I are now in ``REV 2"
Processing status
(see http://heasarc.gsfc.nasa.gov/docs/asca/ascarev2.html), thus
in this paper we have used this new revision of the data.
Furthermore, in order to improve the statistics, we have combined
\footnote{This combination was possible for all the sources but a0447-0627,
a0506-3726, a0506-3742 and a0721+7111. For these 4 sources we have used
only the GIS2 data.}
data from the GIS2 and GIS3 instruments as explained below.
Data preparation has been done using version 1.3 of the XSELECT
software package and version 4.0 of FTOOLS (supplied by the HEASARC at
the Goddard Space Flight Center). Good time intervals were selected
applying the ``Standard REV 2 Screening" criteria (as reported in
chapter 5 of the ASCA Data Reduction Guide, rev 2.0), with the only
exception of having used a magnetic cutoff rigidity threshold of 6 GeV
c$^{-1}$ (as done in Paper I). HIGH, MEDIUM and LOW bit rate data were
combined together. Spectral analysis (see below) has been performed
using version 9.0 of the XSPEC software package. We use the detector
Redistribution Matrix Files (RMF) gis2v4\_0.rmf and gis3v4\_0.rmf.
\vfill
\eject
\subsection{``Stacked" spectra}
For each source and for the GIS2 and GIS3 data sets,
total counts (source + background) were extracted from
a circular extraction region of 2 arcmin radius around the source centroid.
Background counts were taken from two
circular uncontaminated regions of 3.5 arcmin radius, close to the source,
or symmetrically located with respect of the center of the image. Source and
background data were
extracted in the ``Pulse Invariant" (PI) energy channels, which have
been corrected for spatial and temporal variations of the detector
gain. The Ancillary Response File (ARF) relative to each source was
created with version 2.72 of the FTOOLS task ASCAARF
\footnote{ The task ASCAARF (which is part of the FTOOLS software
package) is able to produce a position-dependent PSF-corrected
effective area, A(E,x,y,d), of the (XRT + GIS2 or XRT + GIS3) combination. The
inputs of this task are the position {\it x,y} (in detector
coordinates) and the dimension {\it d} of the source's extraction
region. }
at the
location of the individual sources in the detectors.
For each source we have then produced a combined GIS spectrum
(adding GIS2 and GIS3 data) and the corresponding
background and response matrix files, following the
recipe given in the ASCA Data Reduction (rev 2.0) Guide (see section
8.9.2 and 8.9.3 and reference therein).
Finally, we have produced the combined
spectrum of
a) the 20 sources belonging to the Bright Sample and
b) the 40 sources of the Faint Sample.
We note that each object contributes to the stacked counts at most for
6\% in the case of the Faint Sample and at most for 25\% in
the case of the Bright Sample.
In the spectral analysis, being interested in comparing these stacked
spectra with that of the hard CXB, we have considered only the counts
in the 2.0 - 10.0 keV energy range. The total net counts in the Bright
and Faint sample are about 3400 and 2900 respectively. The stacked
spectra were rebinned to give at least 50 total counts per bin.
\subsection{Hardness Ratios}
For each source and for the GIS2 and GIS3 data set, source plus
background counts were extracted in three energy bands: 0.7-2.0 keV
(S band), 2.0-4.0 keV (M band) and 4.0-10.0 keV (H band). The S, M
and H spectral regions were selected so as to have similar statistics
in each band for the majority of the sources. The background counts
have been evaluated by using the two background regions considered
above; first, we have normalized the background counts to the source
extraction region and then we have averaged them. Net counts in S, M
and H have been then obtained for each source by subtracting the
corresponding S,M and H normalized background counts from the total ones.
To combine the GIS2 and GIS3 data for each source and to
compare our results with those expected from simple models, the net counts obtained
must be corrected for the position dependent sensitivity
of the GIS detectors (see the discussion in section 2.3 of Paper I).
In particular we must:
a) define a source extraction region and a reference position in the GIS2 or
GIS3 field;
b) re-normalize the S,M,H net counts from each source to this region (since
sources are detected in different locations of the GIS2 or GIS3 field of view);
c) perform the simulations for simple models by using the effective area of
this region. We will now discuss these points in turn.
As reference region we have used a source extraction region of 2 arcmin
radius at the position x=137, y=116 for the (XRT + GIS2) combination.
Using ASCAARF we have produced the effective area values at the
position of each source detected in the GIS2 (GIS3) detector. Using
these effective area values and through spectral simulations with XSPEC
we have derived the correction factors to be applied to each source in
the S, M and H band.
For each source we have then applied these correction factors to
the net counts of GIS2 and GIS3 separately.
Finally, we
have combined the GIS2 and GIS3 corrected net counts for each source.
In summary, using this procedure we have first re-normalized the GIS2
and GIS3 data for each source to the reference region separately and
then combined them.
We note that the applied method is very similar to the
``flat field" procedure normally used in the analysis of optical
imaging and spectroscopic data.
Using the corrected net counts in the S, M and H band we have then computed
for each source two ``hardness ratios" defined in the following way:
$$HR1 = {M-S \over M+S}\ \ \ \ \ \ \ HR2 = {H-M \over H+M}$$
The 68\% error bars on HR1 and HR2 have been obtained via Monte Carlo simulations
using the total counts, the background counts and the
correction factors relative to each source.
Similarly, HR1 and HR2 values expected from simple spectral models (see
section 4) have been obtained with XSPEC using the effective area of
the reference region.
\section{Results}
\subsection{The Hard Energy Range and the CXB}
In this section we discuss the hard (2-10 keV) X-ray average spectrum of the
present sample and we compare it with that of the CXB; to this end we
use the HR2 values and the ``stacked" spectra introduced in section 3.
In figure 1, for all sources, we plot the HR2 value versus the GIS2 CCR;
the filled squares represent the sources detected with a
signal to noise ratio greater than 4.0 while the open squares represent
the sources detected with a signal to noise ratio between 3.5 and 4.0.
The HR2 values are then compared with those expected from a non
absorbed power-law model with energy spectral index $\alpha_E$ ($f_X
\propto E^{-\alpha_E}$) ranging from $-$1.0 to 2.0.
Figure 1 clearly shows a broadening of the HR2 distribution going to
fainter CCR; furthermore a flattening of the mean spectrum with
decreasing fluxes is evident. A similar broadening and flattening of
the HR2 distribution is still detected if only the 40 sources (18
Bright and 22 Faint) detected with a signal to noise ratio greater than
4.0 are used, showing that this result is not due to the sources near
the detection threshold limit.
It is worth noting the presence of many sources which seem to be
characterized by a very flat 2-10 keV spectrum with $\alpha_E \leq 0.5$
and of a number of sources with ``inverted" spectra (i.e. $\alpha_E \leq
0.0$). This is particularly evident in the Faint Sample where about
half of the sources seem to be described by $\alpha_E \leq 0.5$ and
about 30\% by ``inverted" spectra. These latter objects could represent
a new population of very hard serendipitous sources or, alternatively,
a population of very absorbed sources as expected from the CXB
synthesis models based on the AGN Unification Scheme.
We have checked whether the observed hardening of the mean spectral
index can be attributed to a spectral bias in the source's selection.
Since sources with the same flux but different spectrum will deposit a
different number of counts in the detector, it is evident that, as one
approaches the flux limit of a survey, sources with a {\it favourable}
spectrum will be detected (and thus included in the sample), while
sources with an {\it unfavourable} spectrum will become increasingly
under-represented (see Zamorani et al., 1988 for a discussion of this
effect). However, in the case of the ASCA/GIS, this selection effect
favours the detection of steep sources, thus giving further support to
the reality of our findings
\footnote{As an example, in a 50.000 s observation, a source with 2-10
keV flux = $5 \times 10^{-13}$ ergs cm$^{-2}$ s$^{-1}\ $ characterized by an absorbed
power-law spectrum with $\alpha_E = 2.0$ and $N_H = 3\times10^{20}$
will deposit (at the reference position and inside a region of 2 arcmin
radius) $\sim 300$ (2-10) keV counts while a source with the same flux
but power-law spectrum with $\alpha_E = 0.0$ and same $N_H$ will
deposit $\sim 180$ counts.}.
We note that a population of very hard X-ray sources has been recently
suggested also by Giommi et al., 1998 in order to explain the spectral
properties of the faint BeppoSAX sources and to reconcile the 2-10 keV
LogN($>$S)--LogS with the 5-10 keV LogN($>$S)--LogS obtained from
BeppoSAX data.
To further investigate the flattening of the source's mean spectral
index we have used the ``stacked" spectra introduced in section 3.1.
In Table 1 we report the results of the power-law fits to the stacked
spectra of the Faint and the Bright sample; the unfolded spectra
of the two samples are shown in figure 2 (Bright sample: open squares;
Faint sample: filled squares).
As can be seen from the $\chi^{2}_{\nu}$ values reported in Table 1
and from the spectra shown in figure 2, a simple power-law spectral
model represents a good description of the stacked spectra of the two
samples in the 2-10 keV energy range. The $N_H$ values corresponding
to the mean line of sight Galactic absorption for the two independent
samples have been used in the fits. However, given the energy range of
interest (2.0 - 10.0 keV), the spectra are not significantly affected
by the Galactic $N_H$ value. Consistent results are obtained if we use
either the lowest or highest ($0.7 \times 10^{20}$ cm$^{-2}$ or $9.06
\times 10^{20}$ cm$^{-2}$ respectively) Galactic $N_H$ value sampled in
the present survey.
It is worth noting that the unfolded spectra of the Faint and the Bright
sample
reported in figure 2 does not show any compelling evidence of emission lines.
A strong emission line should be expected if, for example, sources with
a strong Iron emission line at 6.4 keV, contributing
in a substantial way to the CXB, were strongly clustered at some particular
redshift. This subject has been already discussed by Gilli et
al., 1999, reaching the conclusion that the maximum
contribution of the Iron line to the CXB is less than few percent.
The unfolded spectra shown in figure 2 confirm their results.
In Figure 3 the results obtained here are compared with those obtained
using data from other satellites or from other ASCA medium-deep survey
programs. The best-fit energy index of the Bright sample ($<\alpha_E>
= 0.87 \pm 0.08$) is in good agreement with the mean spectral
properties of the objects in the Ginga and HEAO1 A-2 sample and with
the mean spectral properties of the broad line AGNs detected by ROSAT
(Almaini et al., 1996; Romero-Colmenero et al., 1996). The Faint
sample is best described by $<\alpha_E> = 0.36 \pm 0.14$; this is
consistent with other ASCA results (Ueda et al., 1998) and with the
spectra of the CXB in the 2-10 keV energy range (Marshall et al., 1980;
Gendreau et al., 1995). For the purpose of figure 3 we have
used a count rate to flux conversion factor adeguate for a power law
spectral model with $\alpha_E = 0.36$ (Faint sample) or $\alpha_E =
0.87$ (Bright sample).
We have evaluated the influence that the sources with the hardest energy
distribution have on the combined spectra. If we exclude the 6 sources
with HR2$>$0.2 (5 from the Faint sample and 1 from the Bright sample)
we find that the combined spectra are still significantly different,
being described by a power-law model with $<\alpha_E> = 0.53 \pm 0.14$
and $<\alpha_E> = 1.04 \pm 0.10$ respectively.
Finally, in the case of the Bright sample two sources contribute about 35\%
of the total counts; if we exclude these two objects the
remaining stacked spectrum is described by a power-law model with
$<\alpha_E> = 1.00 \pm 0.16$, showing that the inclusion or the
exclusion of these two objects do not change
any af our results. Note that in the case of the Faint sample each
object contributes to the stacked counts at most for 6\% or less.
These results clearly show that:
a) we have detected a flattening of the source's mean spectral
properties toward fainter fluxes and
b) we are beginning to detect those X-ray sources having a combined
X-ray spectra, consistent with that of the 2 - 10 keV CXB.
\subsection{The Broad Band Spectral Properties and the AGN Unification Scheme}
In this section we do not intend to derive specific spectral properties
and/or parameters for each source; the limited statistics and the
complexity of AGNs broad band X-ray spectra (see Mushotzky, Done
and Pounds, 1993 for a review of the subject) prevent us from doing so.
Rather we regard this sample as representative of the hard X-ray sky
and we try to investigate if the currently popular CXB synthesis models
based on the AGNs Unification Scheme can describe the overall spectral
properties of the ASCA sample as inferred from the hardness ratios.
According to the AGNs Unification model for the
synthesis of the CXB, a population of unabsorbed (Type 1:
$N_{H} \mathrel{<\kern-1.0em\lower0.9ex\hbox{$\sim$}} 10^{22}$ cm$^{-2}$) and absorbed (Type 2: $N_{H} \mathrel{>\kern-1.0em\lower0.9ex\hbox{$\sim$}} 10^{22}$
cm$^{-2}$) AGNs can reproduce the shape and intensity of the
CXB from several keV to $\sim$ 100 keV (see Madau, Ghisellini and
Fabian, 1994; Comastri et al., 1995; Maiolino et al., 1998b).
Because $\sim 90\%$ of the ASCA sources in this sample
\footnote{Up to now 12 sources have been spectroscopically identified.
The optical breakdown is the following: 1 star, 2 cluster of galaxies,
7 Broad Line AGNs and 2 Narrow Line AGNs. However we stress that this
small sample of identified objects is probably not representative of
the whole population.}
are expected to be AGNs (see Paper I), in the following we will consider this sample
as being well approximated by a population of (Type 1 + Type 2) AGNs
with flux above $\sim 1 \times 10^{-13}$ ergs cm$^{-2}$ s$^{-1}\ $.
In figure 4 we show (open squares) the position of the sources in the
``Hardness Ratio" diagram for the Faint (panel a) and Bright (panel b)
sample; in figure 4c we have reported the 12 sources spectroscopically
identified so far (see footnote 8). Also reported in figure 4a,b,c
(solid lines) are the loci expected from X-ray spectra described by an
absorbed power law model. The energy index of the power law ranges
from $-$1.0 up to 2.0, while the absorbing column densities ranges
from $10^{20}$ cm$^{-2}$ up to $10^{24}$ cm$^{-2}$ (see figure 4d);
the absorption has been assumed to be at zero redshift (i.e. the
sources are supposed to be in the local universe).
Figure 4 clearly shows how the ``Hardness Ratio" diagram, combined with
spectral simulations, can be used to obtain information on the sources'
X-ray spectral properties, as well as the power of using two hardness
ratios to investigate the broad band spectral properties of the
sources. If only one ratio is known, say HR2 for example, a source
with HR2 $\sim 0-0.1$ could be described by an absorbed power law
with either energy index of 0.0 and $N_H$ = 10$^{20}$ cm$^{-2}$ or energy
index of $\sim $ 2.0 and $N_H$ = 10$^{23}$ cm$^{-2}$. The use of both
HR1 and HR2 allows the ambiguity to be solved.
We note that the hardness ratios computed as described in section 3.2
have not be corrected for the different Galactic absorbing column
density along the line of sight of each source. For the present sample
this ranges from $10^{20}$ cm$^{-2}$ to $9 \times 10^{20}$ cm$^{-2}$;
figure 4 clearly show that this correction is insignificant.
One of the most striking features of figure 4 is the large spread in HR1
and HR2 displayed by the ASCA sources and the departure from the loci
of absorbed, single power law spectra. This implies that the broad
band (0.7 - 10 keV) spectrum of the sources is more complex than the
simple model of an absorbed power law. In particular the ASCA sources
located on the left side of the line representing power laws with $N_H
\sim$ 10$^{20}$ cm$^{-2}$ (see figure 4d) are not explained within the
absorbed power-law model. A similar result is obtained even if the
absorbing material is assumed to be at higher redshift e.g. z = 1 or 2.
We note that the object in figure 4b (or 4c) characterized by HR1
$\sim$ 0.0 and HR2 $\sim 0.71$ (a1800+6638; its X-ray flux is $\sim
6.5 \times 10^{-13}$ ergs cm$^{-2}$ s$^{-1}\ $) is spectroscopically identified with the
Seyfert 2 galaxy NGC 6552. The X-ray spectrum of this object (Reynolds
et al., 1994 and Fukazawa et al., 1994) is consistent with a model
composed of a narrow Gaussian line plus an empirical ``leaky-absorber"
continuum; the latter is composed by an absorbed power-law and a non
absorbed power law having a common photon index. Fukazawa et al.
(1994) found that the NGC 6552 spectrum requires a photon index of
$\sim 1.4$, an absorbing column density of $\sim 6\times 10^{23}$
cm$^{-2}$, an uncovered fraction of $\sim 2\%$ and a narrow Gaussian
line (consistent with the $K_{\alpha}$ iron emission line at 6.4 keV)
having an equivalent width of $\sim $ 0.9 keV.
That the simple absorbed power law model is unable to explain the
scatter in the hardness ratios is not surprising given the
observational evidence that the broad band X-ray spectra of Type 1
and, even more, of Type 2 AGN are complex and affected by several
parameters like, for instance, the viewing angle and the torus
thickness (see e.g. Turner et al., 1997; Turner et al., 1998;
Maiolino et al., 1998a; Bassani et al., 1999). In particular, for Type
2 AGNs we could have, as a function of the absorbing column density,
one of the following three cases:
1) $10^{22} \mathrel{<\kern-1.0em\lower0.9ex\hbox{$\sim$}} N_H \mathrel{<\kern-1.0em\lower0.9ex\hbox{$\sim$}} 5\times 10^{23}$ cm$^{-2}$. The hard X-ray
continuum above a few keV is dominated by the directly-viewed
component making the source nucleus visible to the observer and the
column density measurable; the reflected/scattered component is
starting to become relevant in the soft energy range. In this case
the observed spectrum is that of an absorbed Type 1 AGNs with some
extra flux at low energies;
2) $5\times 10^{23} \mathrel{<\kern-1.0em\lower0.9ex\hbox{$\sim$}} N_H \mathrel{<\kern-1.0em\lower0.9ex\hbox{$\sim$}} 10^{25}$ cm$^{-2}$. In this case
both the directly-viewed component and the reflected/scattered
component are observed and the resulting spectrum becomes
very complex.
3) $N_H \mathrel{>\kern-1.0em\lower0.9ex\hbox{$\sim$}} 10^{25}$ cm$^{-2}$. The torus is very thick. The
continuum source is blocked from direct view up to several tens of
keV and the observed spectrum is dominated by the scattered/reflected component.
In this case the observed continuum is that of a Type 1 AGNs (in the
case of scattering by warm material near the nucleus) or that of a
Compton reflected spectrum (in the case of cold reflection from the
torus).
Given this spectral complexity we have tested the Hardness Ratios
plot against the following simplified AGNs spectral model composed of:
a) an absorbed power-law spectrum. This component represents the continuum
source, i.e. for an absorbing column density of about $10^{20}$ cm$^{-2}$
this represents the ``zero" order continua of a Type 1 object;
b) a non absorbed power-law component, characterized by the same energy
spectral index of the absorbed power-law and by a
normalization in the range 1-10\% of that of the absorbed power-law
component. This component represents the scattered fraction (by warm
material) of the nuclear emission along the line of sight;
theoretical and observational evidence (see Turner et al., 1997 and
reference therein) suggest that this scattered fraction is in the range
1-10\%;
c) a narrow emission line at 6.4 keV, having an equivalent width of
230 eV for low values of $N_H$ ($\sim 10^{20}$ cm$^{-2}$). This
component represents the mean equivalent width of the Fe K$_{\alpha}$
emission line in Seyfert 1 galaxies as determined by Nandra et al.,
1997 using ASCA data.
In summary, the AGN spectral model used is:
$$S(E) = E^{-\alpha_{E}} e^{-\sigma_{E} \times N_{H}} +
K E^{-\alpha_{E}} + Fe_{6.4 keV}$$
where E is the photon energy, $\alpha_{E}$ is the energy spectral
index, $\sigma_{E}$ is the energy dependent absorbing cross-section,
$N_{H}$ is the absorbing column density along the line of sight to the
nucleus, K is the scattered fraction and $Fe_{6.4 keV}$ is the Iron
narrow emission line at 6.4 keV.
The above simplified AGNs spectral model is equivalent to the model
used, by Fukazawa et al., 1994 in the case of NGC 6552 and, by several
authors to describe the ``first order" X-ray spectra of
Seyfert 2 galaxies (see e.g. Turner et al., 1997).
We have tested this model as a function of $\alpha_{E}$, the absorbing
column density $N_{H}$ and the scattered fraction K. Figure 5 shows,
as an example, the results of the spectral simulations in the case of
local sources ($z\sim 0.0$) with $\alpha_{E} = 1$ and K = 0.01 (filled
triangles) or K = 0.1 (filled circles). The input spectral models for
the case K = 0.01 and Log$N_H$ = 20, 22, 23, 24 and 24.5 are also shown.
We would like to stress that using this simplified AGNs spectral model
we can go, in a continuous way, from a ``first order" Type 1 spectra to a
``first order" Type 2 spectra only by changing the $N_H$ parameter. In
this respect we note that the equivalent width of the Iron narrow
emission line at 6.4 keV (which, as we said, is fixed to 230 eV for
$N_H \sim 10^{20}$ cm$^{-2}$) is $\sim 2$ keV when $N_H \sim
3\times 10^{24}$ cm$^{-2}$; this latter value is very similar to that
measured in many Seyfert 2 galaxies characterized by a similar value of
the absorbing column density (Bassani et al., 1999).
As anticipated at the beginning of this section some qualitative
conclusions can be drawn from the results reported in Figure 5. First
of all it appears that in the context of this simplified AGN
spectral model, and for very high $N_H$ values, we are able to
explain the hardness ratios of the objects located on the left side of
the dashed line representing unabsorbed power laws. Second, the
results reported in Figure 5 seem to indicate that many of the latter
objects could be characterized by a very high absorbing column density;
if this indication is confirmed by deeper investigation with XMM
and AXAF, then the number of Compton ``thick" systems could be
significantly higher than previously estimated. In the meantime we
note that this result is consistent with recent findings obtained by
Maiolino et al., 1998a, Bassani et al., 1999 and
Risaliti, Maiolino and Salvati, 1999 by studying the fraction
of Compton ``thick" systems in a sample of Seyfert 2 galaxies observed
with BeppoSAX and/or with ASCA.
Finally, to investigate the Hardness Ratio plot {\it in a model
independent way}, we have used a comparison sample of nearby (z$<$1)
Seyfert 1 and Seyfert 2 galaxies pointed at with ASCA. This sample was
selected from a data set of about 300 ASCA pointings that were treated
in the same manner (see section 2) and that are used to extend the
survey presented in Cagnoni, Della Ceca and Maccacaro (1998). Amongst
this restricted ASCA data set we have considered those observations
pointed on nearby Seyfert 1 and Seyfert 2 galaxies; furthermore, we
have considered only the Seyfert 2 galaxies also reported in Table 1 of
Bassani et al.,1999 in order to have an uniform data set on their
Compton thickness.
The comparison sample is composed of 13
Seyfert 1, 15 Compton thin Seyfert 2 and 7 Compton thick Seyfert 2
\footnote{The Seyfert 1 are: NGC 1097, MKN 1040, RE 1034+39, MKN 231, MKN 205,
3C 445, I ZW 1, 2E1615+0611, MKN 478, EXO055620-3820.2, MKN 507, NGC 985, TON 180.
The Compton thin Seyfert 2 are: NGC 3079, NGC 4258, NGC 3147, NGC 5252,
MKN 463e, NGC 1808, NGC 2992, NGC 1365, NGC 6251, MKN 273, IRAS05189-2524, NGC 1672,
PKS B1319-164, MKN 348, NGC 7172.
The Compton thick Seyfert 2 are: MKN 3, NGC 6240, NGC 1667, NGC 4968, NGC 1386,
MKN 477, NGC 5135.}.
The targets were analyzed in the same way as the
serendipitous sources, combining the GIS2 and GIS3 data
(see section 3.2).
The results are shown in figure 6. The Seyfert 1 galaxies are strongly
clustered around the locii representing low $N_H$ values; on the
contrary the Seyfert 2 galaxies are characterized by a very large
spread in their HR1, HR2 values. In particular many of the Seyfert 2
are located on the left side of the line representing power laws with
$N_H \sim 10^{20}$ cm$^{-2}$; about a half of these sources have been
classified as Compton thick system from Bassani et al., 1999. We are
tempted to suggest that also the other half are Compton thick objects,
whose nature is still unrevealed because of the presently poor data
quality. XMM and/or AXAF observations are needed to confirm this
suggestion. However, whatever their properties (Compton Thin or Thick)
are, the comparison of figures 6 and figures 4a,b strongly suggest
that at least some of the serendipitous ASCA sources located on the
left side of the dashed line connecting the $N_H=10^{20}$ values could
be Type 2 AGNs.
\section{Summary and Conclusion}
In this paper we have used ASCA GIS2 and GIS3 data for a complete
and well defined sample of 60 hard (2-10 keV) selected sources to study
the spectral properties of X-ray sources down to a flux limit
of $\sim 10^{-13}$ ergs cm$^{-2}$ s$^{-1}\ $.
To investigate if the spectral properties of the sources depend on
their brightness we have defined two subsamples according to the
``corrected'' 2-10 keV count rate; the Bright sample is defined
by the 20 sources with CCR $\geq 3.9\times 10^{-3}$ cts s$^{-1}\ $ , while the
Faint sample is defined by the 40 fainter sources.
The dividing line of $3.9 \times 10^{-3}$ cts s$^{-1}\ $
corrispond to unabsorbed 2-10 keV fluxes in the range $\sim
5.4\times 10^{-13}$ to $\sim 3.1\times 10^{-13}$ ergs cm$^{-2}$ s$^{-1}\ $ for a source
described by a power law model with energy spectral index between 0.0
and 2.0, respectively.
The main results of this investigation are the following:
a) the average (2-10 keV) source's spectrum hardens towards fainter
fluxes. The ``stacked" spectra of the sources in the Bright sample
is well described by a power-law model with
an energy spectral index of $<\alpha_E> = 0.87\pm 0.08$, while the
``stacked" spectra of the sources in the Faint sample
require $<\alpha_E> = 0.36 \pm 0.14$; this means that we are beginning
to detect those
sources having a combined X-ray spectrum consistent with that of the
2-10 keV CXB.
b) the hardness ratios analysis shows that this flattening is due to
the appearance of sources with very hard spectra in the Faint sample.
About a half of the sources in the latter sample
require $\alpha_E \mathrel{<\kern-1.0em\lower0.9ex\hbox{$\sim$}} 0.5$ while only 10\%
of the brighter sources are consistent with an energy spectral index so
flat. Furthermore about 30\% of the sources in the Faint sample
seem to be characterized by ``inverted'' ($\alpha_E \mathrel{<\kern-1.0em\lower0.9ex\hbox{$\sim$}}
0.0$) 2-10 keV X-ray spectra. These latter objects could represent a
new population of very hard serendipitous sources or, alternatively, a
population of very absorbed sources as expected from the CXB synthesis
models based on the AGN Unification Scheme (see below).
c) a simple absorbed power-law model is unable to explain the broad
band (0.7 - 10 keV) spectral properties of the sources, as inferred
from the Hardness Ratios diagram.
X-ray spectral
models based on the AGNs Unification Scheme seem to be able to explain
the overall spectral properties of this sample. This seems also
suggested by
the comparison of our results with those obtained using a
sample of nearby and well known Seyfert 1 and Seyfert 2 galaxies
observed with ASCA.
Acknowledgments
We are grateful to L.Bassani for stimulating discussions on the
comparison sample of Seyfert galaxies, A.Wolter for a carefull reading
of the manuscript and the anonymous referee for useful comments. G.C.
acknowledge financial support from the ``Fondazione CARIPLO". This
work received partial financial support from the Italian Ministry for
University and Research (MURST) under grant Cofin98-02-32. This
research has made use of the NASA/IPAC extragalactic database (NED),
which is operated by the Jet Propulsion Laboratory, Caltech, under
contract with the National Aeronautics and Space Administration. We
thank all the members of the ASCA team who operate the satellite and
maintain the software data analysis and the archive.
\newpage
\begin{center}
Table 1\\
\vskip 0.5truecm
\begin{tabular}{ccccrr} \hline \hline
Sample & Objects & Net counts & $\alpha_E$ & $N_{H Gal}$ & $\chi^2_{\nu}$/$\nu$\\ \cline{1-6} \hline
& & $(2-10)$ keV & & $(10^{20}cm^{-2})$ &\\ \hline
Faint & 40 & $\sim 2900$ & $0.36\pm0.14$ & 2.75 & 1.09/105\\
Bright & 20 & $\sim 3400$ & $0.87\pm0.08$ & 3.66 & 1.06/82\\ \hline
\end{tabular}
\end {center}
|
\section*{Figure Captions}
\begin{description}
\item[Fig. 1]
Top view of a random distribution of air-cylinders.
The circles refer to the air-cylinders (not to scale).
A straight line source is
placed at the origin.
\item[Fig. 2] The scattering function versus frequency.
\item[Fig. 3] Transmission versus frequency in terms of $ka$.
\item[Fig. 4] Left column: Phase diagram for the two-dimensional phase
vector defined in the text. Right column: Spatial distribution of acoustic
energy (arbitrary scale).
\item[Fig. 5] `Order parameters' versus frequency.
\end{description}
\end{document}
|
\section{INTRODUCTION}
\label{sec:intro} \vspace*{0.2cm}
Since Bekenstein and Hawking found that the black hole entropy is
proportional to the event horizon area by comparing black hole
physics with thermodynamics and from the discovery of the black
hole evaporation \cite{Bekenstein72}-\cite{Bekenstein74}, many
efforts have been devoted to study the statistical origin of the
black hole entropy. Especially, the idea to relate the entropy of
the black hole to quantum excitations of the black hole has
attracted many attentions \cite{Hooft85} -\cite{Mann96}. The
thermodynamical entropy of the black hole is related to the
covariant Euclidean free energy $F^E[g, \beta]=\beta^{-1}W[g,
\beta]$\cite{Frolov98}, where $\beta$ is the inverse temperature.
The function $W[g, \beta]$ is given on Euclidean manifolds with
the period $\beta$ in the Euclidean time $\tau$. We can calculate
the free energy $F^E$ by the conical singularities method. This
procedure was consistently carried out for the studies of the
static black holes and the rotating charged Kerr black
hole\cite{Solodukhin951} \cite{Solodukhin952} \cite{Fursaev95}
\cite{Fursaev96} \cite{Mann96}. On the other hand, the canonical
statistical-mechanical entropy can be derived from the free energy
$F^C$ of a system\cite{Frolov98}, where $F^C$ can be defined in
term of the one particle spectrum. One of the ways to calculate
$F^C$ is the ``brick wall" model (BWM) proposed by 't Hooft
\cite{Hooft85}. He argued that the black hole entropy is
identified with the statistical-mechanical entropy arising from a
thermal bath of quantum fields propagating outside the horizon. In
this model, in order to eliminate divergence which appears due to
the infinite growth of the density of states closed to the
horizon, 't Hooft introduces a ``brick wall" cutoff: a fixed
boundary $\Sigma_h$ near the event horizon within the quantum
field does not propagate and the Dirichlet boundary condition was
imposed on the boundary, i. e., wave function $\phi=0$ for
$r=r({\Sigma_h})$. Later, J. G. Demers, R, Lafrance and R. C.
Myers \cite{Demers95} pointed out that the Dirichlet condition can
be removed if we use the Pauli-Villars regulated theory. The BWM
has been successfully used in studies of the
statistical-mechanical entropy for many black holes
\cite{Hooft85}, \cite{Ghosh94}-\cite{Ho98}, \cite{Belgiorno96},
\cite{Jing97}.
Recently, Frolov and Fursaev\cite{Frolov98} reviewed the studies
of the relation between the thermodynamic entropy and the
statistical-mechanical entropy of the black holes. They shown
that for the general static black holes the covariant Euclidean
free energy $F^E$ and the statistical-mechanical free energy $F^C$
are equivalent when ones use the ultraviolet regularization method
\cite{Frolov98}.
As static case, the quantum entropy for the stationary
axisymmetric black holes has also been studied by many authors
recently. Mann and Solodukhin\cite{Mann96} investigated the
covariant Euclidean formulation for the Kerr-Newman black hole.
They showed that an Euclidean manifold which was obtained by Wick
rotation of the Kerr-Newman geometry with Killing horizon has a
conical singularity similar to the one which appears in the static
black holes. The one-loop quantum correction to the entropy of the
charged Kerr black hole was calculated by applying the method of
the conical singularities. They found an interesting result that
the logarithmic term of the quantum entropy for the Kerr-Newman
black hole can be written as a constant plus a term proportional
to the charge and so, for the Schwarzschild and the Kerr black
holes, the logarithmic parts in the entropy are exactly equal.
Cognola \cite{Cognola98}, through the Euclidean path integral and
using heat kernel and $\zeta$-function regularization scheme,
studied the one-loop contribution to the entropy for a scalar
field in the Kerr black hole. In the calculation he took an
approximation of the metric, which, after a conformal
transformation, takes a Rindler-like form. He pointed out in
Ref.\cite{Cognola98} that the result is valid also for the
Kerr-Newman black hole. Nevertheless, the result is in contrast
with the corresponding one obtained in Ref.\cite{Mann96}.
In Refs. \cite{Lee96} \cite{Lee961} \cite{Ho96}, by using the 't
Hooft BWM Lee and Kim, Ho, Kim and Park discussed the
statistical-mechanical entropy of some stationary black holes,
such as Kerr black hole, Kerr-Newman black hole, and Kaluza-Klein
black hole. The results showed that the entropies can be expressed
as $k\frac{A_H}{\varepsilon ^2}$, where $A_H$ is the area of the
event horizon, and $\varepsilon=\int^{r_H+h}_{r_H}dr\sqrt{g_{rr}}$
is the proper distance from the horizon to the $r_H+h$ and the $h$
the cutoff in the radial coordinate near the horizon. However, the
logarithmically divergent term of the statistical-mechanical
entropy for the four-dimensional stationary axisymmetric black
hole space-time was not be investigated.
Although much attention has been paid on the study of the quantum
entropy of the stationary axisymmetric black holes, the relation
between the statistical-mechanical entropy and the thermodynamical
entropy for the rotating black holes has not been investigated
yet\cite{Frolov98}. The aim of this paper is to obtained an
expression of the general statistical-mechanical entropy for the
general four-dimensional non-extreme stationary axisymmetric
black hole by using the BWM and the Pauli-Villars regularization
scheme, and then make a comparison of the statistical-mechanical
entropy obtained by using BWM and the thermodynamical entropy by
the conical singularity method for some well-known stationary
axisymmetric black holes.
The paper is organized as follows: In sec.2, the general
stationary axisymmetric black hole is introduced and some
properties of the black hole that it is necessary to understand
the thermodynamics of quantum field are studied. In sec. 3,
making use of 't Hooft's BWM \cite{Solodukhin97} we deduce a
formula of the statistical-mechanical entropy for the general
non-extreme stationary axisymmetric black hole. In the last
section, the statistical-mechanical entropies for the Kerr-Newman
and the Einstein-Maxwell dilaton-axion (EMDA) black holes are
studied by using the formula. Then the results are compared with
the entropy obtained by the conical singularity method. Finally,
we end with some conclusions.
\section{the space-time of the general non-extreme \\
Stationary axisymmetric black hole}
\vspace*{0.4cm}
In Boyer-Lindquist coordinates the most general line element for a
stationary axisymmetric black hole in four-dimensional space-time
can be expressed as
\begin{equation}
ds^2=g_{tt}dt^2+g_{rr}dr^2+g_{t \varphi}dtd\varphi+g_{\theta
\theta}d\theta ^2+g_{\varphi \varphi}d\varphi ^2,
\label{metric0} \end{equation}
where $g_{tt}$, $g_{rr}$, $g_{t\varphi}$, $g_{\theta \theta}$ and
$g_{\varphi \varphi}$ are functions of the coordinates $r$ and
$\theta$ only. Because the space-time (\ref{metric0}) is
stationary and axisymmetric one it exists a stationary Killing
vector field $\xi^{\mu}=(1, 0, 0, 0)$ and an axial Killing field
$\Psi^{\mu}=(0,0,0,1)$\cite{Carmeli}. By taking a liner
combination of $\xi^\mu$ and $\Psi^\mu$ we obtain a new Killing
field
\begin{equation}
l^\mu=\xi^\mu+\Omega_{H}\Psi^\mu,\label{lll}
\end{equation}
which is normal to the horizon of the black hole. In Eq.
(\ref{lll}) the constant $\Omega _{H}$ is called the angular
velocity of the event horizon. An interesting feature of the
black hole worthy of note is that the
norm of the Killing field $l^\mu$ is zero on the horizon since the
horizon is a null surface and vector $\l^\mu$ is normal to the
horizon. That is to say, the black hole horizon is a surface where
the Killing field $l^\mu$ is null. Substituting $l^\mu$ into the
formula of the surface gravity \cite{Wald} $
\kappa^2=-\frac{1}{2}\l^{\mu
;\nu}l_{\mu ;\nu}, $ we obtain
\begin{eqnarray}
\kappa&=&\frac{-1}{2}\lim_{r\rightarrow r_H}\left
(\sqrt{\frac{-1}{g_{rr}\left(g_{tt}-
\frac{g_{t\varphi}^2}{g_{\varphi \varphi}}\right)}}\frac{d}{dr}
\left(g_{tt}-\frac{g_{t\varphi}^2}{g_{\varphi \varphi}}\right)
\right)\nonumber \\
&=&\frac{2\pi}{\beta_H}
, \label{surface}
\end{eqnarray}
where $r_H$ represents the outermost event horizon, $1/\beta_H$ is
the Hawking temperature, and here and hereafter the metric
signature is taken as $(-, +, +, +)$. We know that the event
horizon is a null hypersurface determined by
\begin{equation}
g^{\mu \nu}\frac{\partial H}{\partial x^\mu}\frac{\partial H}{\partial
x^\nu}=0. \label{hs1}
\end{equation}
For the stationary axisymmetric black hole (\ref{metric0}) the
function $H$ is a function of $r$ and $\theta$ only.
Substituting the metric (\ref{metric0}) into Eq. (\ref{hs1}) and
discussing carefully we find
\begin{equation}
\frac{1}{g^{tt}(r_H)}=\left(g_{tt}-\frac{g_{t\varphi}^2}{g_{\varphi\varphi}}
\right)_{r_H}=0. \label{nullcon}
\end{equation}
Solutions of which determine the location of the event horizons.
From Eq. (\ref{nullcon}) we know that for a non-extreme
stationary axisymmetric black hole $1/g^{tt}$ can be expressed as
\begin{equation}
\left(g_{tt}-\frac{g_{t\varphi}^2}{g_{\varphi\varphi}}
\right)\equiv G_1(r,\theta)(r-r_H), \label{asum}
\end{equation}
where $G_1(r,\theta)$ is a regular function in the region $ \infty>r\geq
r_H$ and its value is nonzero on the outermost event horizon
$ r_H$. On the other hand, since $\kappa=constant$
and $\frac{1}{g^{tt}}=0$ on the event horizon $r=r_H$, we find from
Eq.(\ref{surface}) that $g^{rr}$ must take the following form
\begin{equation}
g^{rr}\equiv G_2(r,\theta)(r-r_H),\label{asum1}
\end{equation}
where $G_2(r,\theta)$ is a well-define function in the region $
\infty>r\geq r_H$ and is nonzero on the horizon $r_H$ too. Making
use of Eqs. (\ref{asum}) and (\ref{asum1}), we obtain
\begin{equation}
g_{rr}\left(g_{tt}-\frac{g_{t\varphi}^2}{g_{\varphi\varphi}}
\right)=\frac{G_1(r, \theta)}{G_2(r, \theta)}
\equiv -f(r,\theta), \label{relation}
\end{equation}
where the $f(r, \theta)$ is a constant or a regular function on the
outermost event horizon and outside the horizon.
\section{The statistical-mechanical entropy of general \\
non-extreme stationary axisymmetric black hole} \vspace*{0.5cm}
We now try to find an expression of the statistical-mechanical
entropy due to the minimally coupled quantum scalar fields in a
general four-dimensional stationary axisymmetric black hole. We
first seek the total number of modes with energy less than $E$ by
using Klein-Gordon equation, and then calculate a free energy. The
statistical-mechanical entropy of the black hole is obtained by
the variation of the free energy with respect to inverse
temperature and setting $\beta =\beta_H$.
Using WKB approximation with
\begin{equation}
\phi =exp[-iEt+im\varphi+iW(r,\theta)], \label{phi}
\end{equation}
and substituting the metric (\ref{metric0}) into the
Klein-Gordon equation of the scalar field $\phi$ with mass $\mu$
and nonminimal $\xi R \phi^2$ ($R$ is the scalar curvature of the
spacetime) coupling
\begin{equation}
\frac{1}{\sqrt{-g}}\partial _\mu(\sqrt{-g}g^{\mu\nu}\partial _\nu
\phi)-(\mu ^2+\xi R)\phi=0, \label{kg}
\end{equation}
we find \cite{Mann92}
\begin{equation}
p_r^2=\frac{1}{g^{rr}}
[-g^{tt}E^2+2g^{t\varphi}Em-g^{\varphi\varphi}
m^2-g^{\theta\theta}p_\theta^2-(\mu ^2+\xi R)],\label{W}
\end{equation}
where $p_r\equiv \partial _r W(r, \theta)$ and $p_\theta \equiv
\partial_\theta W(r, \theta)$.
If the scalar curvature $R$ takes a nonzero value at the horizon
then this region can be approximated by some sort of constant
curvature space. However, the exact results for such a black hole
showed that the mass parameter in the solution enters only in the
combination $(\mu^2-R/6)$\cite{Solodukhin97}\cite{Birrell82}.
Therefore, inserting the covariant metric into Eq.(\ref{W}) we
arrive at
\begin{equation}
p_r^2=-\frac{g_{rr}g_{\varphi\varphi}}
{g_{tt}g_{\varphi\varphi}-g_{t\varphi}^2} \left[(E-\Omega
m)^2+\left(g_{tt}-\frac{g_{t\varphi}^2}{g_{\varphi\varphi}}
\right)\left(\frac{m^2}{g_{\varphi\varphi}}+\frac{p_\theta^2}
{g_{\theta\theta}}+M^2(r,\theta)\right)\right],\label{pr}
\end{equation}
where $\Omega \equiv -\frac{g_{t \varphi }}{g_{\varphi \varphi}}$
and $M^2(r,\theta)\equiv \mu^2-(\frac{1}{6}-\xi)R$. In this paper
our discussion is restricted to study minimally coupled ($\xi=0$)
scalar fields. We know from Eq. (\ref{pr}) that $W(r,\theta)$ can
be expressed as
\begin{equation}
W(r,\theta)=\pm \int^r
\sqrt{\frac{-g_{rr}g_{\varphi\varphi}}{g_{tt}g_{\varphi\varphi}-
g_{t\varphi }^2}} K(r, \theta) dr+c(\theta),
\end{equation}
where
\begin{equation}
K(r, \theta)=\sqrt{(E-\Omega m)^2+\left(g_{tt}-
\frac{g_{t\varphi}^2}{g_{\varphi\varphi}}
\right)\left(\frac{m^2}{g_{\varphi\varphi}}+\frac{p_\theta^2}
{g_{\theta\theta}}+M^2(r, \theta)\right)}.\label{KK}
\end{equation}
Therefore, in phase space the number of the modes with $E$, $m$
and $p_{\theta}$ is shown by\cite{Padmanabhan86}
\begin{equation}
n(E, m, p_{\theta})=\frac{1}{\pi} \int d\theta
\int^{r_E}_{r_H+h}\sqrt{\frac{-g_{rr}g_{\varphi
\varphi}}{g_{tt}g_{\varphi \varphi}-g_{t\varphi}^2}}K(r,\theta)dr.
\end{equation}
Zhao and Gui \cite{Zhao83} pointed out that ``a physical space"
must be dragged by the gravitational field with a azimuth angular
velocity $\Omega_{H}$ in the stationary axisymmetric space-time
(\ref{metric0}). Apparently, a quantum scalar field in thermal
equilibrium at temperature $1/\beta$ in the stationary
axisymmetric black hole must be dragged too. Therefore, it is
rational to assume that the scalar field is rotating with angular
velocity $\Omega_0=\Omega_{H} $. For such an equilibrium ensemble
of states of the scalar field, the free energy is given by
\begin{eqnarray}
\beta F&=& \int dm \int dp_{\theta}\int dn(E, m, p_{\theta})ln
\left[ 1-e^{-\beta (E-\Omega_0 m)}\right] \nonumber \\ &=&\int dm
\int dp_{\theta}\int dn(E+\Omega_0 m, m, p_{\theta})ln \left(
1-e^{-\beta E}\right) \nonumber \\ &=&-\beta \int dm \int
dp_{\theta}\int \frac{n(E+\Omega_0 m, m, p_{\theta})} {e^{\beta
E}-1} dE \nonumber \\ &=&-\beta \int \frac{n(E)}{e^{\beta E}-1} dE
, \label{f1}
\end{eqnarray}
with
\begin{eqnarray}
n(E)&=&\int dm \int dp_{\theta}\int n(E+\Omega_0 m, m, p_{\theta})
\nonumber \\ &=&\frac{1}{3\pi}\int d\theta
\int^{r_E}_{r_H+h}\frac{dr \sqrt{g_4} }
{\left[\left(g_{tt}-\frac{g_{t\varphi }^2}{g_{\varphi
\varphi}}\right)\left(1+\frac{g_{\varphi
\varphi^2(\Omega-\Omega_0)^2}}{g_{tt}g_{\varphi \varphi}-g_{t
\varphi}^2} \right)\right]^2} \left[E^2 \right. \nonumber \\ & &
\left. +\left(g_{tt}-\frac{g_{t\varphi}^2}{g_{\varphi \varphi}}
\right)\left(1+\frac{g_{\varphi
\varphi^2(\Omega-\Omega_0)^2}}{g_{tt}g_{\varphi \varphi}-g_{t
\varphi}^2} \right)M^2(r,\theta)\right]^{\frac{3}{2}}. \label{nE}
\end{eqnarray}
where the function $n(E)$ presents the total number of the modes
with energy less than $E$. The integrations of the $m$ and
$p_{\theta}$ in the Eq.(\ref{f1}) are taken only over the value
for which the square root in Eq.(\ref{KK}) exists.
Taking the integration of the $r$ in the Eq.(\ref{nE}) for the
case $\Omega_0=\Omega_H$ we have
\begin{eqnarray}
n(E)&=&-\frac{1}{2\pi}\int d\theta \left \{\sqrt{g_{\theta
\theta}g_{\varphi \varphi}}\left[\frac{2}{3}\left(\frac{E
\beta_H}{4\pi} \right)^3C(r,\theta)+M^2(r,\theta) \left(\frac{E
\beta_H}{4\pi}\right)\right] ln \left(\frac{E^2}{E^2_{min}}\right)
\right\}_{r_H}
\nonumber \\
& &-\frac{1}{3\pi}\left(\frac{\beta_H}{4\pi}\right)\int d\theta
\left\{\sqrt{g_{\theta
\theta}g_{\varphi\varphi}}M^2(r,\theta)
\left(E-\frac{E^3}{E^2_{min}}\right)\right\}_{r_H}
, \label{n0}
\end{eqnarray}
where
\begin{eqnarray}
C(r,\theta)&=&\frac{\partial ^2g^{rr}}{\partial
r^2}+\frac{3}{2}\frac{\partial g^{rr}}{\partial r}\frac{\partial
\ln f}{\partial
r}-\frac{2\pi}{\beta_H\sqrt{f}}\left(\frac{1}{g_{\theta
\theta}}\frac{\partial g_{\theta \theta}}{\partial r}+
\frac{1}{g_{\varphi \varphi}}\frac{\partial g_{\varphi
\varphi}}{\partial r}\right)-\frac{2g_{\varphi
\varphi}}{f}\left[\frac{\partial}{\partial r}\left(\frac{g_{t
\varphi}}{g_{\varphi \varphi}}\right)\right]^2,\nonumber\\
E^2_{min}&=&-M^2(r_H, \theta)\left(g_{tt}-\frac{g_{t\varphi^2}}{g_{\varphi
\varphi}}\right)_{\Sigma_h}, \nonumber \\
\bar{K}^2&=&E^2+\left(g_{tt}-\frac{g_{t\varphi}^2}{g_{\varphi
\varphi}}\right)_{\Sigma_h}M^2(r_H, \theta),
\end{eqnarray}
here and hereafter $f\equiv f(r,\theta)$ which is defined by Eq.
(\ref{relation}).
Now let us use the Pauli-Villars regularization scheme
\cite{Demers95} by introducing five regulator fields $\{\phi_i,
i=1,...,5\}$ of different statistics with masses
$\{\mu_i,i=1,...,5\}$ dependent on the UV cutoff \cite{Demers95}.
If we rewrite the original scalar field $\phi=\phi_0$ and
$\mu=\mu_0$, then these fields satisfy $\Sigma^5_{i=0}\triangle
_i=0$ and $\Sigma^5_{i=0}\triangle _i \mu^2_i=0$, where
$\triangle_0=\triangle_3=\triangle_4=+1$ for the commuting fields
and $\triangle_1=\triangle_2=\triangle_5=-1$ for the anticommuting
fields. Since each of the fields makes a contribution to the free
energy of Eq.(\ref{f1}), the total free energy can be expressed as
\begin{eqnarray}
\beta \bar{F}=\sum^5_{i=0}\triangle _i \beta F_i. \label{f2}
\end{eqnarray}
Substituting Eq.(\ref{f1}) into (\ref{n0}) and then taking the
integration over E we find
\begin{eqnarray}
\bar{F}&=&\frac{-1}{48}\frac{\beta_H}{\beta ^2}\int
d\theta\left\{\sqrt{g_{\theta \theta}g_{\varphi
\varphi}}\right\}_{r_H} \sum^5_{i=0}\triangle _i
M_i^2(r_H,\theta)lnM_i^2(r_H, \theta)
-\frac{1}{2880}\frac{\beta_H^3}{\beta^4} \int
d\theta\left\{\sqrt{g_{\theta \theta}g_{\varphi \varphi}} \right.
\nonumber \\ & &\times \left. \left[\frac{\partial
^2g^{rr}}{\partial r^2}+ \frac{3}{2}\frac{\partial
g^{rr}}{\partial r}\frac{\partial \ln f}{\partial
r}-\frac{2\pi}{\beta_H \sqrt{f}}\left(\frac{1}{g_{\theta
\theta}}\frac{\partial g_{\theta \theta}}{\partial r}+
\frac{1}{g_{\varphi \varphi}}\frac{\partial g_{\varphi
\varphi}}{\partial r}\right) -\frac{2g_{\varphi
\varphi}}{f}\left[\frac{\partial}{\partial r}\left(\frac{g_{t
\varphi}}{g_{\varphi \varphi}}\right)\right]^2
\right]\right\}_{r_H} \nonumber \\ & &\times
\sum^5_{i=0}\triangle
_ilnM_i^2(r_H,\theta), \label{f-0}
\end{eqnarray}
where $M_i^2(r_H, \theta)=\mu_i^2-\frac{1}{6}R$. Then the total
statistical-mechanical entropy at the Hawking temperature
$\frac{1}{\beta}=\frac{1}{\beta_H}$ is given by
\begin{eqnarray}
S&=&\left[\beta^2 \frac{\partial \bar{F}}{\partial
\beta}\right]_{\beta=\beta_H}\nonumber \\ &=&\frac{1}{48\pi}\int
d\theta d\varphi\left(\sqrt{g_{\theta \theta}g_{\varphi
\varphi}}\right)_{r_H} \sum^5_{i=0}\triangle _i
M_i^2(r_H,\theta)lnM_i^2(r_H, \theta) +\left\{\frac{1}{32\times
45\pi}\int d\theta d\varphi \sqrt{g_{\theta \theta}g_{\varphi
\varphi}} \right. \nonumber \\ & & \left. \times
\left[\frac{\partial ^2g^{rr}}{\partial r^2} + \frac{ 3 }{ 2
}\frac{ \partial g^{rr} }{
\partial r } \frac{\partial \ln f}{\partial r}-\frac{2\pi}{\beta_H
\sqrt{f}}\left(\frac{1}{g_{\theta \theta}}\frac{\partial g_{\theta
\theta}}{\partial r}+ \frac{1}{g_{\varphi \varphi}}\frac{\partial
g_{\varphi \varphi}}{\partial r}\right)-\frac{2 g_{\varphi
\varphi}}{f}\left[\frac{\partial}{\partial r}\left(\frac{g_{t
\varphi}}{g_{\varphi
\varphi}}\right)\right]^2\right]\right\}_{r_H}\nonumber \\ &
&\times
\sum^5_{i=0}\triangle _ilnM_i^2(r_H,
\theta). \label{SM}
\end{eqnarray}
Using the assumption that the scalar curvature $R$ at the horizon
is much smaller than each $\mu_i$ and noting that the area of the
event horizon is given by $A_{\Sigma}= \int d\varphi \int
d\theta\left\{\sqrt{g_{\theta \theta}g_{\varphi
\varphi}}\right\}_{r_H}$, we obtain at last the following
expression of the statistical-mechanical entropy
\begin{eqnarray}
S&=&\frac{A_{\Sigma}}{48\pi}\sum^5_{i=0} \triangle _i
\mu_i^2ln\mu_i^2+\left\{-\frac{1}{6\times 48 \pi} \int d\theta
d\varphi \left(R\sqrt{g_{\theta \theta}g_{\varphi
\varphi}}\right)_{r_H}+\frac{1}{32\times 45\pi}\int d\theta
d\varphi \right.\nonumber \\ & & \times \left.
\left[\sqrt{g_{\theta \theta}g_{\varphi \varphi}}\left(
\frac{\partial ^2g^{rr}}{\partial r^2}+ \frac{3}{2}\frac{\partial
g^{rr}}{\partial r}\frac{\partial \ln f}{\partial
r}-\frac{2\pi}{\beta_H \sqrt{f}}\left(\frac{1} {g_{\theta
\theta}}\frac{\partial g_{\theta \theta}}{\partial r}+
\frac{1}{g_{\varphi \varphi}}\frac{\partial g_{\varphi
\varphi}}{\partial r}\right) -\frac{2g_{\varphi \varphi}}{f}
\right. \right. \right. \nonumber \\ & & \left. \left. \left.
\times \left[\frac{\partial}{\partial r}\left(\frac{g_{t
\varphi}}{g_{\varphi
\varphi}}\right)\right]^2\right)\right]_{r_H}\right\}
\sum^5_{i=0}\triangle _iln\mu_i^2. \label{smu}
\end{eqnarray}
which is valid for the general non-extreme stationary
axisymmetric black holes which the metric can be expressed as
(\ref{metric0}) in the Boyer-Lindquist coordinates and their
signature is $(-, +, +, +)$. For the black holes with signature
$(+, -, -, -)$ a corresponding formula can be obtained by replaced
the $\beta_H$ with $-\beta_H$ in the Eq.(\ref{smu}).
\section{discussion and summary}
In this section, let begin discussion with the study of the
statistical-mechanical entropy of the Kerr-Newman black hole and
EMDA black hole by using the formula (\ref{smu}).
\vspace*{0.4cm}
{\bf (A) The entropy of the Kerr-Newman black
hole}
\vspace*{0.4cm}
In Boyer-Lindquist coordinates, the metric of the Kerr-Newman
black hole \cite{Kerr63}\cite{Newman65} takes the form
\begin{eqnarray}
g_{tt}&=&-\frac{\bigtriangleup -a^2sin^2\theta }{\rho ^2}, \ \ \ \
g_{t\varphi}=-\frac{asin^2\theta
(r^2+a^2-\bigtriangleup)}{\rho ^2},\nonumber \\
g_{rr}&=&\frac{\rho ^2}{\bigtriangleup}, \ \ \ \ g_{\theta \theta
}=\rho ^2, \ \ \ \ g_{\varphi \varphi
}=\left(\frac{(r^2+a^2)^2-\bigtriangleup a^2sin^2\theta }{\rho
^2}\right)sin^2 \theta,\label{knm}
\end{eqnarray}
with
\begin{equation}
\rho^2=r^2+a^2cos^2\theta,\ \ \ \ \bigtriangleup=(r-r_+)(r-r_-),
\end{equation}
where $r_+=r_H=M+\sqrt{M^2-Q^2-a^2}$, $r_-=M-\sqrt{M^2-Q^2-a^2}$,
and $M$, $Q$ represent the mass and charge of the black hole,
respectively. Using the metric (\ref{knm}) we get
\begin{eqnarray}
& &\left\{\frac{\partial ^2g^{rr}}{\partial r^2}+
\frac{3}{2}\frac{\partial g^{rr}}{\partial r}\frac{\partial \ln
f}{\partial r}-\frac{2\pi}{\beta_H \sqrt{f}}\left(\frac{1}
{g_{\theta \theta}}\frac{\partial g_{\theta \theta}}{\partial r}+
\frac{1}{g_{\varphi \varphi}}\frac{\partial g_{\varphi
\varphi}}{\partial r}\right) -\frac{2g_{\varphi \varphi}}{f}
\left[\frac{\partial}{\partial r}\left(\frac{g_{t
\varphi}}{g_{\varphi
\varphi}}\right)\right]^2\right\}_{r_+}\nonumber
\\ &=&\frac{16r_+^2[(r_+r_--a^2)-(r_++r_-)r_+]+4[(a^2-r_+r_-)+3(r_
++r_-)r_+]\rho^2}{\rho^6}+\frac{2(a^2-r_+r_-)}{\rho^4}\nonumber
\\ & &+\frac{2a^2(1+cos^2\theta )\rho ^2-8a^2(r_+^2+a^2)cos^2\theta
}{\rho^6}, \nonumber \\ R&=&0. \label{kn1}
\end{eqnarray}
Inserting Eq.(\ref{kn1}) into Eq.(\ref{smu}) and then taking the
integrations of the $\theta$ and $\varphi$ we find that the
statistical-mechanical entropy of the Kerr-Newman black hole is
given by
\begin{eqnarray}
S_{KN}&=&\frac{A_{\Sigma}}{48\pi}\sum^5_{i=0} \triangle _i
\mu_i^2ln\mu_i^2-\frac{1}{90}\left\{1+\frac{3(a^2-r_+
r_-)}{4r_+^2}\left[1+\frac{r_+^2+a^2}{ar_+}Arctan\left(
\frac{a}{r_+}\right)\right]\right\} \sum^5_{i=0}\triangle
_iln\mu_i^2, \nonumber\\ \label{kn2}
\end{eqnarray}
where $A_{\Sigma}=4\pi(r_+^2+a^2)$. Noting $r_+r_--a^2=Q^2$ and
the Pauli-Villars regularization scheme caused a factor
$-\frac{1}{2}$ for the second part in the Eq.(\ref{kn2}), we know
that the statistical-mechanical entropy (\ref{kn2}) coincides
with the Mann-Solodukhin's result \cite{Mann96} which obtained by
using the conical singularity method.
\vspace*{0.4cm}
{\bf (B) The entropy of the Stationary
axisymmetric EMDA black hole} \vspace*{0.4cm}
The stationary axisymmetric EMDA black hole metric (we take the
solution b=0 in Eq.(14) in Ref.(2). The reason we use this
solution is that the solution $b\neq0$ cannot be interpreted
properly as a black hole) is described by\cite{Garcia95}
\begin{eqnarray}
g_{tt}&=&-\frac{\bigtriangleup -a^2sin^2\theta }{\Sigma }, \ \ \ \
g_{t\varphi}=-\frac{asin^2\theta
[(r^2+a^2-2dr)-\bigtriangleup]}{\Sigma},\nonumber \\
g_{rr}&=&\frac{\Sigma}{\bigtriangleup}, \ \ \ \ g_{\theta \theta
}=\Sigma, \ \ \ \ g_{\varphi \varphi
}=\left(\frac{(r^2+a^2-2dr)^2-\bigtriangleup a^2sin^2\theta
}{\Sigma}\right)sin^2 \theta, \label{EMDA}
\end{eqnarray}
with
\begin{equation}
\Sigma=r^2-2dr+a^2cos^2\theta,\ \ \ \
\bigtriangleup=r^2-2mr+a^2=(r-r_+)(r-r_-),
\end{equation}
where $r_+=m+\sqrt{m^2-a^2}$, $r_-=m-\sqrt{m^2-a^2}$. The mass M,
the angular momentum J, the electric charge Q,
and the magnetic charge P of the black hole are respectively given by
\begin{equation}
M=m-d,\ \ \ \ J=a(m-d),\ \ \ \ Q=\sqrt{2\omega d(d-m)}, \ \ \ \
P=g.
\end{equation}
By using the metric (\ref{EMDA}) we obtain
\begin{eqnarray}
& &\left\{\frac{\partial ^2g^{rr}}{\partial r^2}+
\frac{3}{2}\frac{\partial g^{rr}}{\partial r}\frac{\partial \ln
f}{\partial r}-\frac{2\pi}{\beta_H \sqrt{f}}\left(\frac{1}
{g_{\theta \theta}}\frac{\partial g_{\theta \theta}}{\partial r}+
\frac{1}{g_{\varphi \varphi}}\frac{\partial g_{\varphi
\varphi}}{\partial r}\right) -\frac{2g_{\varphi \varphi}}{f}
\left[\frac{\partial}{\partial r}\left(\frac{g_{t
\varphi}}{g_{\varphi
\varphi}}\right)\right]^2\right\}_{r_+}\nonumber
\\ &=&\frac{16r_+^2[(2d-r_+-r_-)r_+]+4d(8r_+-3d)
(r_+^2-2dr_++a^2)}{\Sigma^3}\nonumber \\ & &+
\frac{4[(3r_+-2d)(r_++r_--2d)-d^2]}{\Sigma^2}
+\frac{2d(r_++r_--2d)(\Sigma-2dr_+)}{\Sigma^3}\nonumber
\\ & & +\frac{2a^2(1+cos^2\theta )\Sigma-8a^2(r_+^2+a^2-2dr_+)cos^2\theta
}{\Sigma^3},\nonumber \\
R&=&\frac{2a^2d^2sin^2\theta}{\Sigma^3}.
\label{EMDA1}
\end{eqnarray}
Substituting Eq.(\ref{EMDA1}) into Eq.(\ref{smu}) and then taking
the integration of the $\theta$ and $\varphi$ we find that the
statistical-mechanical entropy of the EMDA black hole is
\begin{eqnarray}
S&=&\frac{A_{\Sigma}}{48\pi}\sum^5_{i=0} \triangle _i
\mu_i^2ln\mu_i^2-\frac{1}{90}\left\{1+\frac{9d^2}{8r_+^2-16dr_+}
+\frac{9d\{3a^2d+(r_+^2-2dr_+)[\frac{4}{3}(r_++r_-)-d]\}}{16(r_+^2-2dr_+)^2}
\right. \nonumber \\ & & \left. \times
\left[1+\frac{r_+^2+a^2-2dr_+} {a\sqrt{r_+^2-2dr_+}}Arctan\left(
\frac{a}{\sqrt{r_+^2-2dr_+}}\right) \right] \right\}
\sum^5_{i=0}\triangle _iln\mu_i^2, \label{EMDA2}
\end{eqnarray}
where $A_{\Sigma}=4\pi(r_+^2+a^2-2dr_+)$. In order to compare the
entropy (\ref{EMDA2}) with the thermodynamical entropy obtained
by the covariant Euclidean formulation, we \cite{Jing99}
calculated the thermodynamical entropy of the EMDA black hole by
using the conical singularity method of the Ref.\cite{Mann96}. We
also find that the results obtained by the two methods are
equivalent.
From Eq.(\ref{kn2}) or Eq.(\ref{EMDA2}) we find the same result
for the Kerr black hole (setting $ Q=0$ in Eq. (\ref{kn2}) or
$d=0$ in Eq.(\ref{EMDA2})) as that Mann and Solodukhin found in
Ref.\cite{Mann96}, i. e., the quantum entropy does not depend on
the rotation parameter $a$ and coincides with the quantum entropy
of the Schwarzschild black hole. We think the reason is that the
quantum entropy is mainly caused by quantum scalar fields near the
event horizon and in the region the scalar fields are co-rotating
with the black hole.
To summary, by using BWM and with Pauli-Villars regularization
scheme, we investigate the statistical-mechanical entropy arising
from the minimally coupled quantum scalar fields rotating with the
angular velocity $\Omega_0$ in the general four-dimensional
non-extreme stationary axisymmetric black hole space-time. An
expression of the statistical-mechanical entropy is obtained for
the case $\Omega_0=\Omega_H$. The Kerr-Newman black hole and the
EMDA black hole are studied. It is shown that the
statistical-mechanical entropy obtained by using the formula
(\ref{smu}) and the quantum thermodynamical entropy derived from
the covariant Euclidean formulation (by using the conical
singularity method) are equivalent for the Kerr-Newman and the
EMDA black holes. The result fills in the gaps mentioned in
Ref.\cite{Frolov98} that the relation between the canonical and
covariant Euclidean formulations in the rotating black hole has
not been investigated. The study may provide us with a better
understanding of the relationship between the different entropies
to the black holes.
|
\section{Introduction}
The microscopic origin of the sliding friction of thin films
adsorbed on a metal surface has recently attracted considerable
interest~\cite{bowden,persson}. The availability of refined
experimental methods~\cite{krim1,krim2,holzapfel} makes it now
feasible to investigate the fundamental processes that
contribute to the sliding friction on an atomic scale. On a
metal surface, frictional energy is dissipated via two main
channels: excitation of low-energy electron-hole pairs in the
substrate~\cite{SH81,Sols82,P91,PV95,sokoloff95,liebsch97,liebsch97b}
and excitation of phonons within the
overlayer~\cite{cieplak94,cieplak94b,PN96,TS97}. Phonon
excitation within the metal may also occur; however, this
contribution usually has only a minor influence on the lateral
sliding of adsorbed films.
Detailed experimental information on the friction of Xe and Kr
layers on noble metal surfaces was recently obtained by
Krim {\it et al.}~\cite{krim1,krim2} who performed quartz-crystal
microbalance (QCM) measurements for a variety of coverages. In this
technique, the frequency shift and broadening of the resonant line
provide information on the number of adsorbed atoms and on their
slippage during the lateral oscillation of the microbalance.
The data were analyzed by various theoretical groups that,
surprisingly, arrived at contradictory conclusions: For Kr on Au(111),
Cieplak {\it et al.}~\cite{cieplak94,cieplak94b} argued that the
observed friction
is caused by lattice vibrations within the overlayer and that
substrate-induced energy dissipation plays a negligible role. Dominant
phonon damping was also inferred for Xe on Ag(111)
by Tomassone {\it et al.}~\cite{TS97}.
On the other hand, Persson {\it et al.}~\cite{PN96} argued that the
phonon-associated friction for a full Xe monolayer on Ag(100) is
small and that the measured friction is mainly of electronic origin.
These opposite interpretations are puzzling since all three theoretical
studies were based on molecular dynamics simulations for
similar models of the rare gas/metal adsorption system. In addition,
Liebsch~\cite{liebsch97,liebsch97b} performed dynamical surface response
calculations within the time-dependent density functional approach
to determine the electronic friction of Xe atoms on Ag. The lateral
component of the friction coefficient was shown to agree qualitatively
with the experimental value for a full monolayer, suggesting that
phonon processes play a minor role in the compressed phase.
In view of these fundamental theoretical discrepancies, we
decided to carry out molecular dynamics simulations for Xe on
Ag, employing essentially the same overall model as previous
authors~\cite{cieplak94,cieplak94b,PN96,TS97}. Key input
parameters in these simulations are, in particular, the
corrugation amplitude of the Xe/Ag interaction potential and the
electronic friction of individual Xe atoms sliding parallel to
the substrate surface. Since at present the corrugation
amplitude is known from independent experimental data only with
very appreciable uncertainty, a detailed understanding of how
its magnitude influences the final result is in our view
crucial. In contrast to the earlier molecular dynamics studies,
therefore, we do not choose a specific set of input model
parameters but investigate the net sliding friction within a
wide range of these parameters. In addition, we consider the
dependence on the crystallographic orientation by studying the
friction of a Xe layer adsorbed on Ag(111) and Ag(100), in order
to check the importance of the lateral symmetry of the substrate
potential. By carrying out such a systematic investigation we
essentially cover the full range of overlayer/substrate models
considered in the previous treatments.
A striking feature of the Xe/Ag quartz-crystal microbalance data by
Krim {\it et al.}~\cite{krim1,krim2} is the strong variation of the
net sliding friction with Xe coverage. To facilitate the discussion,
we show in Fig.~1 the measured slip time $\tau$ as a function
of Xe coverage~\cite{krim2} on Ag(111). The friction coefficient
$\bar\eta$ is given by the inverse of the slip time. According
to surface resistivity measurements for Xe on Ag~\cite{holzapfel},
the electronic friction coefficient depends weakly
on coverage. It is evident, therefore, that the rapid coverage
dependence observed in the QCM data, in particular near completion
of a monolayer, is caused primarily by
phonons generated within the overlayer. This is not surprising
since the probability of exciting phonons varies greatly with
inter-adatom spacing. With respect to this aspect of the data,
there is little dispute among the previous simulations studies.
\begin{figure}
\includegraphics[width=8.cm, clip=true]{fig1_fric.eps}
\narrowtext
\caption{Measured slip time $\tau$ for Xe on Ag(111),
as a function of coverage, after Ref.~4.
In this paper we put forward arguments, based on extensive
simulations, to show that the region roughly between the
dotted lines can be understood as due to phonon originated friction,
while above the upper dotted line a friction mechanism of electronic
origin seems to prevail.}
\end{figure}
The important question, rather, concerns the relative weights of
electronic and phononic dissipation channels, especially at
coverages near a full monolayer where the slip time reaches its
maximum, i.e., where the sliding friction of the incommensurate
overlayer is smallest. Our results show that the effective
friction parameter $\bar\eta$ may be approximately written as
\begin{equation}
\bar\eta = \eta_{el} + \eta_{ph}\ .
\end{equation}
\noindent Here, $\eta_{el}=\eta_\Vert$ is the parallel component
of the single-atom electronic friction coefficient and the
intra-adsorbate phonon contribution is roughly given by
\,$\eta_{ph}=c\,u_0^2$, where $u_0$ is the corrugation amplitude
of the Xe/Ag potential. A quadratic variation with $u_0$ agrees
with the analytical and numerical results of Cieplak {\it et
al.}~\cite{cieplak94,cieplak94b}. The coefficient $c$ depends
weakly on the crystal face and on the single-atom friction
parameters $\eta_\Vert$ and $\eta_\perp$, but varies strongly
with coverage. Evidently, according to the above expression, the
measured sliding friction at any given coverage can be
reproduced using several combinations of the parameters $u_0$
and $\eta_{el}$. For instance, by reducing the single-site
electronic friction $\eta_\Vert$, one is able to match the data
if the substrate potential is assumed to be sufficiently
strongly corrugated. Conversely, if the electronic friction is
chosen large enough, little additional phonon damping is
required to reproduce the same net $\bar\eta$. Constraints on
the range of acceptable parameter values are implied by the
observed variation of the slip time with coverage, in
particular, by the steep slope near completion of the first
monolayer.
The molecular dynamics simulations for Xe on Ag, which we describe
below, suggest that the characteristic coverage dependence of the
observed slip time and its overall magnitude cannot be understood
only in terms of phonon dissipation. Although there is considerable
uncertainty both in the calculated and measured friction
coefficients, the most likely scenario is the presence of a roughly
coverage-independent electronic contribution of about \,$\tau\approx
2\,{\rm ns}$\, ($\eta_\Vert\approx0.5\,{\rm ns}^{-1}$) and an additional,
strongly coverage-dependent phonon friction, which can reduce the slip
time to about \,$\tau\approx1\,{\rm ns}$ (the net sliding friction then
increases approximately to \,$\bar\eta\approx1\,{\rm ns}^{-1}$).
Thus, roughly speaking, we associate the slip time in the region between
the horizontal dotted lines in Fig.~1 with phonon processes, while the
portion above the upper line is related to electronic energy dissipation.
Remarkably, the electronic friction deduced from this analysis is in
qualitative agreement with independent surface resistivity data for
Xe on Ag~\cite{holzapfel} and with theoretical estimates of this
dissipation channel~\cite{liebsch97,liebsch97b}. Thus, the friction of a full
monolayer is essentially of electronic origin, while the rapid
increase of the slip time before reaching the complete monolayer is
due to excitation of phonons within the Xe overlayer.
This paper is organized as follows: Section II summarizes the
theoretical results available for the electronic friction of Xe atoms
on a metal surface. Section III presents the model employed in the
molecular dynamics simulations. In particular, we specify the
interaction potentials and the friction felt by a single Xe atom.
Various technical aspects of the calculation of the
phonon friction are also discussed. Section IV contains the results,
with special emphasis on the variation of the net sliding friction
with the corrugation of the substrate potential and its dependence on
the electronic friction. Our conclusions are given in Section V.
\section{Electronic Friction}
We consider first the electronic contributions to the sliding friction
of individual Xe atoms above a metal surface. In a physisorption system,
these may arise due to the van der Waals attraction and the Pauli
repulsion. In the case of Xe, some dissipation might also be associated
with the formation of a weak covalent bond.
It is well known that the instantaneous mutual polarization between
neutral species gives rise to the long-range van der Waals attraction.
If an atom moves relative to a metal, the induced surface charge lags
behind, causing kinetic energy to be dissipated. Thus, as pointed out
long ago by various authors~\cite{SH81,Sols82}, the excitation of
low-frequency electron-hole pairs in the surface region contributes
to the friction of neutral atoms. If the adatom velocity is small
compared to the Fermi velocity of the metal electrons, the friction
coefficient may be expressed as \cite{Sols82,PV95}
\begin{equation}
\eta_i = -{e^2\over m}\,\lim_{\omega\rightarrow0}\,{1\over\omega}
\int\!d^3r\!\int\!d^3r' \, V_i({\bf r})
\, {\rm Im}\,\chi({\bf r,r'},\omega)\, V_i({\bf r'}).
\label{etai}
\end{equation}
The index \,$i=x,z$\, refers to parallel or perpendicular motion of
the atom, \,$\chi$\, is the many-body response function of the metal,
and $m$ is the adatom mass. The effective interaction
\,$V_i({\bf r})$\, is approximately given by
\begin{equation}
V_i({\bf r}) = {\alpha(0)\over 2}\,\nabla_i\left(\nabla
{1\over\vert{\bf r - d}\vert}\right)^2 \ .\label{Vi}
\end{equation}
Here, \,$\alpha(0)$\, is the static polarizability of the adatom and
\,${\bf d}=(0,0,d)$\, is the location relative to the jellium edge.
Since \,$V_i({\bf r})$\, does not satisfy the Laplace equation,
it can only be approximately represented in terms of a superposition
of evanescent plane waves of the form \cite{PV95}
\begin{equation}
V_i({\bf r}) = \sum_{\bf q_\Vert}\ V_i({\bf q_\Vert})
\ e^{i{\bf q_\Vert}\cdot{\bf r_\Vert} + qz} \ .
\end{equation}
With this expansion we can write $\eta_i$ as
\begin{equation}
\eta_i={e^2\over m}\ {1\over2\pi }\ \lim_{\omega\rightarrow0}\,
{1\over\omega}
\sum_{\bf q_\Vert}\ \vert V_i({\bf q_\Vert})\vert^2
\ q \ {\rm Im}\,g(q,\omega) \ , \label{img}
\end{equation}
where the surface response function $g(q,\omega)$ is defined as
\begin{equation}
g(q,\omega) = \int dz\ e^{qz}\ n_1(z,q,\omega)
\end{equation}
and \,$n_1(z,q,\omega)$\, is the electronic surface density induced
by an external potential given by
\,$-{2\pi\over q}\,e^{i\vec q_\vert\cdot\vec r_\Vert +qz}$.
In the case of Xe atoms physisorbed on Ag(111) at a temperature
$T=77.4$~K, the distance $d=2.4$~\AA. The static polarizability
of Xe is \,$\alpha(0)=4.0$~\AA$^3$. Using the density functional
results for \,$n_1(z,q,\omega)$\, and \,$g(q,\omega)$, we find
\,$\eta_\Vert\sim 0.34$~ns$^{-1}$\, and
\,$\eta_\perp\sim1.74$~ns$^{-1}$ \cite{liebsch97,liebsch97b}.
We note that the theoretical value of $\eta_\Vert$ given above
presumably represents a slight overestimate since the Xe atom is
not completely outside the range of the electronic density profile
as implicitely assumed in (\ref{img}). Moreover, as a result of
the hybridization between the Ag $s$ and $d$ electrons, the surface
polarizability of real Ag may be slightly smaller than that of
the corresponding jellium model~\cite{inglesfield}. We also point
out that, according to (\ref{Vi}), the friction coefficient varies
asymptotically like \,$1/d^{10}$, i.e., it is rather sensitive to
the precise location of the adatom above the metal surface.
The friction coefficient arising from the van der Waals
attraction derived above is significantly larger than the
contribution due to the Pauli repulsion which was estimated by
Persson~\cite{P91} to be about \,$\eta_\Vert\sim
0.06$~ns$^{-1}$. In addition, for Xe there may exist some
friction due to chemical effects resulting from the broadening
of the Xe $6s$ level. Estimates of this mechanism
yield~\cite{P91} \,$\eta_\Vert\sim 0.15$~ns$^{-1}$.
On the basis of these theoretical calculations and estimates, we
conclude that the total electronic contribution to the lateral
friction of single Xe atoms on Ag is approximately
\,$0.5\ldots0.6$~ns$^{-1}$. At full monolayer coverage, this
friction is presumably slightly smaller because of the weakening of
the Xe/Ag bonds. An electronic contribution of this magnitude is in
excellent agreement with the surface resistivity measurements for
Xe on Ag~\cite{holzapfel}.
According to our molecular dynamics simulations at $T=77.4$~K,
the distance of Xe atoms normal to the surface varies about
0.1~\AA\ around the equilibrium distance \,$d=2.4$~\AA. For the
asymptotic van der Waals attraction, such a variation implies a
variation of $\eta_\Vert$ roughly between 0.2 and 0.5~ns$^{-1}$.
For the Pauli and covalent-bond contributions the variation is
exponential due to wave function overlap, with an exponent
determined by the work function. This leads to a variation of
$\eta_\Vert$ of about 20 \%. Thus, for some Xe atoms, the
electronic friction is larger, for others smaller than the
equilibrium value \,$0.5\ldots0.6$~ns$^{-1}$\, quoted above.
Because of this uncertainty the simulations are carried out for
a sufficiently wide range of electronic friction coefficients.
\section{Model for Phonon Friction}
In this section we focus on the evaluation of the friction induced by
the excitation of phonons within the adsorbed layer of Xe atoms. The
single-atom frictional properties are assumed to be known and are
used as inputs in the molecular dynamics simulation. We specify first
the intra-adsorbate and adsorbate--substrate potentials and then discuss
details concerning the molecular dynamics simulations.
\subsection{Potentials}
As in previous work~\cite{cieplak94,cieplak94b,PN96,TS97}, the interaction
between Xe atoms is expressed as a sum of Lennard-Jones pair
potentials
\begin{equation}
v(r) = \epsilon\,\left[\left(r_0\over r\right)^{12} -
2\left(r_0\over r\right)^6 \right]\ ,
\end{equation}
\noindent
where $\epsilon=19$ meV is the well depth and $r_0=4.54$ \AA\ is
the inter-particle spacing at the potential minimum.
The total Xe-Xe interaction potential then takes the form
\begin{equation}
V = {1\over2}\sum_{i\neq j} v({\bf r}_i -{\bf r}_j)\ .
\end{equation}
\noindent
The adsorbate-substrate interaction potential is assumed to be
given by
\begin{eqnarray}
U(\bf r) &=& E_0\,\left[ e^{-2\alpha(z-z_0)}
- 2 e^{- \alpha(z-z_0)} \right] \nonumber\\
& & + \ u_0\, e^{-\alpha'(z-z_0)}\, u(x,y) \ , \label{U}
\end{eqnarray}
\noindent
where for Ag(100)
\begin{equation}
u(x,y) = 2 + \cos(kx) + \cos(ky) \ ,
\end{equation}
\noindent
whereas, in the case of Ag(111),
\begin{eqnarray}
u(x,y) &=& 1.5 +\cos\Big(2ky/\sqrt3\Big)
+\cos\Big(k(x+y/\sqrt3)\Big)\nonumber\\
&& +\cos\Big(k(x-y/\sqrt3)\Big)\ .
\end{eqnarray}
\noindent
Here, $k=2\pi/a$ and $a=2.89$ \AA\ is the spacing between neighboring
Ag atoms. The unit vectors of the substrate lattice are $(a,0), \ (0,a)$
for Ag(100) and $(a,0), \ (a/2,a\sqrt3/2)$ for Ag(111). The binding
energy for Xe on Ag is $E_0=0.23$~eV. The decay constant $\alpha$ may
be deduced from the frequency of the perpendicular vibration~\cite{PN96},
i.e., $\alpha=0.72$~\AA$^{-1}$. The decay constant $\alpha'$ of the
Fourier component should in principle be larger than $\alpha$.
However, this difference is not very important and we choose
$\alpha'=\alpha$. The corrugation factor $u(x,y)$ is assumed to
vanish at hollow adsorption sites. The potential barrier between
neighboring hollow sites is $2u_0$ for the (100) face and $0.5u_0$
for the (111) face. The total barrier between hollow sites across
the substrate lattice sites is $4u_0$ for the (100) face and $4.5u_0$
for (111).
The adsorbate--substrate interaction specified in Eq.~(\ref{U}) has
the same form as the one used by Persson and Nitzan~\cite{PN96}.
Cieplak {\it et al.}~\cite{cieplak94,cieplak94b} and Tomassone {\it et
al.}~\cite{TS97}, on the other hand, used the expression derived by
Steele~\cite{steele} for a sum of pair potentials. Since this
potential is much more corrugated than the true interaction between
rare gas atoms and metal surfaces, the coefficient of the
non-vanishing Fourier component was drastically scaled down.
The functional form of this potential also differs from expression
(\ref{U}). However, the precise shape near the potential minimum
should not be crucial, as long as the overall corrugation is the
same.
Since the parameter $u_0$ is decisive for the excitation of phonons
within the overlayer, we do not choose a particular value. Instead, we
calculate the net sliding friction as a function of $u_0$ in order to
illustrate its sensitivity to the adsorbate--substrate potential.
In this manner, we cover the various substrate potentials considered
previously.
\subsection{Simulations} There are several methods of extracting
the friction parameter from the molecular dynamics simulations.
(i)~One can apply a constant lateral force $\bf F$ to all the
adatoms, while keeping the substrate atoms at rest. After an
appropriate number of time steps, the steady-state velocity $\bf
v$ of the overlayer is determined. If the friction is viscous,
$\bf v$ satisfies the linear relation \begin{equation} {\bf F} =
m\,\bar\eta \,{\bf v}\ , \end{equation} where $m$ is the adatom
mass. This procedure was used by Cieplak {\it et
al.}~\cite{cieplak94,cieplak94b} and by Persson and Nitzan~\cite{PN96};\
(ii)~To simulate the quartz-crystal microbalance measurement,
one may let the substrate atoms oscillate laterally and
determine the induced lateral vibration of the adatoms. This
method was also used by Cieplak {\it et al.}~\cite{cieplak94,cieplak94b};\
(iii)~One can apply a constant force to the adatoms up to a
specific point in time and then switch it off. If the friction
is viscous, the decay of the steady-state velocity of the
adsorbate is proportional to $e^{-\bar\eta t}$;\ (iv)~Finally,
one may determine $\bar\eta$ from the thermal equilibrium
autocorrelation function of the center of mass velocity of the
overlayer. The latter two methods were employed by Tomassone
{\it et al.}~\cite{TS97}.
In the present work, we also apply a constant lateral force $\bf F$
to the adatoms and determine the steady-state velocity that is reached
after a sufficiently long time. The simulations are based
on the Langevin equation
\begin{equation}
m\ddot{\bf r}_i+m\eta\dot{\bf r}_i = - {\partial U\over\partial{\bf r}_i}
- {\partial V\over\partial{\bf r}_i}
+ {\bf f}_i + {\bf F} \ ,\label{LE}
\end{equation}
\noindent
where ${\bf f}_i$ is a stochastically fluctuating force describing
the effect of the irregular thermal motion of the substrate on the
$i^{th}$ adatom. The components of ${\bf f}_i$ are related to the
microscopic friction $\eta$ via the fluctuation-dissipation theorem
\begin{equation} \langle f_i^\alpha(t)\,f_j^\beta(0) \rangle = 2mk_{\rm B} T
\eta_{\alpha\beta} \delta_{ij}\delta(t)\ .
\end{equation}
\noindent
The microscopic friction tensor $\eta_{\alpha\beta}$ is assumed to be
diagonal, with independent elements $\eta_\Vert$ and $\eta_\perp$.
To solve Eq.~(\ref{LE}) the time variable is discretized with a
step duration $0.01\,t_0$, where the natural time unit is
\,$t_0=r_0\,(m/\epsilon)^{1/2}= 3.84$~ps. The simulations are
carried out using the integration procedure suggested by Tully
{\it et al.}~\cite{tully}. Typically, thermalization was
reached after $10^2-10^3\,t_0$. The drift friction $\bar\eta$
was derived using the definition \,$\bar\eta=F/(m\langle
v\rangle)$, where the external force $F$ was applied in the
$x$-direction on each adsorbate, resulting in the (time
averaged) drift velocity $\langle v\rangle$. On Ag(100), the
basic unit was taken as a square containing $12\times12$ or
$24\times24$ substrate atoms. Similar unit cells were employed
for Ag(111). These cells were then repeated assuming periodic
boundary conditions.
Except for the form of the substrate potential, the model
outlined above is very similar to the ones used in earlier
simulations~\cite{cieplak94,cieplak94b,PN96,TS97}. The main difference
between the treatments Of Refs.~\cite{cieplak94,cieplak94b,PN96,TS97}
resides in the interpretation of the microscopic friction of a
single Xe adatom. In the work by Cieplak {\it et
al.}~\cite{cieplak94,cieplak94b} this friction plays the role of a
thermostat that allows to establish a constant temperature $T$
within the overlayer. Since the direction of this single-site
friction is chosen orthogonal to the direction of the external
force, it was claimed not to affect the net sliding friction of
the overlayer. Instead, in the work by Persson and
Nitzan~\cite{PN96} this thermostat was given a specific
microscopic origin. In particular, $\eta_\Vert$ coincides with
the lateral friction of individual Xe atoms due to excitation of
electron-hole pairs in the substrate, i.e.,
$\eta_\Vert=\eta_{el}$. According to the theoretical estimates
and the resistivity measurements discussed in Section II, this
coefficient should be about \,$\eta_\Vert\approx 0.5$~ns$^{-1}$.
Finally, in the work by Tomassone {\it et al.}~\cite{TS97}, the
simulations were carried out in the absence of a thermostat or
any other single-site friction forces.
In principle, electronic processes also contribute to the friction of
the perpendicular motion of single adatoms. As pointed out in Section
II, density functional calculations yield
\,$\eta_\perp\approx1.74$~ns$^{-1}$.
However, this contribution is much smaller than the normal friction
induced by phonon excitation, which was estimated by Persson and
Nitzan~\cite{PN96} to be roughly \,$\eta_\perp=260$~ns$^{-1}$.
Because of the present uncertainties in the evaluation of both
$\eta_\Vert$ and $\eta_\perp$, we have carried out simulations for
a whole range of these parameters, in order to illustrate the
sensitivity of the net sliding friction to the microscopic
processes governing the behavior of individual Xe atoms.
\section{Results and Discussion} In this section we first
discuss the sliding friction of a monolayer of Xe atoms on Ag
and then address the coverage dependence. In the final part, we
compare these results with the QCM data. The temperature of
\,$T=77.4$~K, which we adopt, corresponds to the one used in the
measurements by Krim {\it et al.}~\cite{krim1}. We define the
coverage \,$\Theta=1$\, as the fully compressed
monolayer~\cite{dai94} with density $n_a=0.0597$~\AA$^{-2}$. On
Ag(100) this amounts to \,$N=72$\, Xe atoms within the basic
\,$12a\times12a$\, unit cell ($N=288$ within the $24a\times24a$
cell). The uncompressed monolayer has a slightly lower density,
\,$n_a=0.0565$~\AA$^{-2}$, i.e., \,$\Theta=0.94$ ($N=68$\, atoms
within the \,$12a\times12a$\, unit cell). The remaining
parameters required in the simulations, which we report below,
have been specified in the preceding sections.
\subsection{Corrugation dependence of the sliding friction for
an uncompressed Xe layer}
In Fig.~2 we display typical results of the center of mass
velocity, in the direction of the applied force, as a function of
time. The examples are for $N=68$ on the (100) surface and the
constant external force on each of the adsorbates \,$F=0.001\,
\epsilon/r_0$\, is applied after \,$100\,t_0$\, as indicated by
the arrow. The single-atom friction coefficients are
\,$\eta_\Vert= 0$\, and \,$1.3$~ns$^{-1}$, with
\,$\eta_\perp=260$~ns$^{-1}$. The corrugation amplitude is
\,$u_0=0.95$~meV and $1.9$~meV.
\begin{figure}
\includegraphics[width=8.cm, clip=true]{fig2a_fric.eps}
\includegraphics[width=8.cm, clip=true]{fig2b_fric.eps}
\narrowtext
\caption{Center of mass velocity (in units \AA/ps)
of a Xe layer on Ag(100) as a function of time. The coverage
corresponds to an uncompressed monolayer at $\Theta = 0.94$ or
$n_a = 0.0565$ atoms/\AA$^2$. The arrow indicates the time
($t=100\,t_0=384$~ps) at which the external force
\,$F=0.001\,\epsilon/r_0=4.2\times 10^{-3}$~meV/\AA\
is applied. Upper curves are for $\eta_\Vert=0$, lower curves
correspond to \,$\eta_\Vert=1.3$~ns$^{-1}$. (a)~$u_0=0.95$~meV;
(b)~$u_0=1.9$~meV.}
\end{figure}
One of the conclusions of the present work can be derived
directly from this figure.
In Fig.~2a, for a relatively low corrugation amplitude
\,$u_0=0.95$~meV, the effect of the single-site friction
coefficient $\eta_\Vert$ is evident: suppressing it, the center
of mass velocity increases and so do the fluctuations. But the
effect of the suppression is attenuated when the corrugation
amplitude is bigger, as is the case in Fig.~2b for
\,$u_0=1.9$~meV. Thus, the relative importance of phononic or
electronic dissipation channels depends crucially on the
magnitudes of $u_0$ and $\eta_\Vert$.
The validity of the linear response regime can be checked in Fig.~3, where
we show the center of mass velocity $\langle v\rangle$, time averaged over
\,$4000\,t_0$\, after the transient period following the application
of the external force $\bf F$, as a function of $F$. The approximate
linear relation \,$F = m\bar\eta\langle v\rangle$\, is apparent in all
cases, indicating that the system indeed obeys a viscous force law.
The full lines represent weighted linear regressions of the
\,$\langle v\rangle\sim F$\, relation, which we use to derive the net
sliding friction $\bar\eta$.
\begin{figure}
\includegraphics[width=7.cm, clip=true]{fig3_fric.eps}
\narrowtext
\caption{Average center of mass velocity $\langle v\rangle$
(in units of $r_0/t_0$) as a function of external force $F$
(in units of $\epsilon/r_0$) showing the linear viscous regime
for different values of the input parameters $\eta_\Vert$ and $u_0$.
Each point was obtained by time
averaging over a long steady state run and over two independent runs.
Error bars were estimated from the fluctuations of the center of mass
velocity as can be seen in Fig.~2. Straight lines are linear fits
of points accounting for error bars. The values of $\bar\eta$
are as follows (from top to bottom):
\,$0.0023/t_0$, \,$0.0045/t_0$, and \,$0.010/t_0$.}
\end{figure}
Fig.~4 summarizes our results for $\bar\eta$ as a function of the
corrugation amplitude $u_0$ and for different values of the lateral
electronic single-atom friction $\eta_\Vert$.
The overall sliding friction is approximately of the form
\begin{equation}
\bar\eta = \eta_{el} + c\, u_0^2\ , \label{c}
\end{equation}
\noindent
with \,$\eta_{el}=\eta_\Vert$. The intra-adsorbate phonon contribution
\,$\eta_{ph}=c\,u_0^2$\, varies quadratically with the corrugation
amplitude $u_0$ and the coefficient $c$ depends weakly on the
coefficient $\eta_\Vert$ and on the crystal structure of the substrate.
This can be seen by comparing the results shown in Fig.~4a and 4b for
the Ag(100) and Ag(111) surfaces, respectively. Although the overall
variation of $\bar\eta$ is similar for both faces, the coefficient $c$
is slightly larger for Ag(111) than for Ag(100), indicating that $\bar\eta$
is governed by the total barrier height ($4.5\,u_0$\, for (111) vs.
$4\,u_0$\, for (100)), rather than by the corrugation barrier between
neighboring hollow sites ($0.5\,u_0$\, for (111) vs. $2\,u_0$\, for
(100); see previous section). This trend is plausible since, at
monolayer coverage, the Xe atoms are not free to move along paths
between nearest potential minima but are forced across potential maxima.
We note here that, in the \,$\eta_\Vert=0$ case, a quadratic variation of
$\bar\eta$ with $u_0$ was also found by Cieplak {\it et
al.}~\cite{cieplak94,cieplak94b}.
\begin{figure}
\includegraphics[width=7.cm, clip=true]{fig4a_fric.eps}
\includegraphics[width=7.cm, clip=true]{fig4b_fric.eps}
\narrowtext
\caption{Net sliding friction of Xe layer on Ag as
a function of corrugation amplitude $u_0$. The coverage is
$\Theta=0.94$. The results are shown for three values of the lateral
single-atom electronic friction, indicated by the symbols:
\,$\eta_\Vert=\eta_{el}=0$, \,$0.0025\,/t_0$, and \,$0.0050\,/t_0$.
\,$\eta_\perp=1.0\,/t_0$ in all cases. (a) (100) surface; (b) (111) surface.}
\end{figure}
In order to demonstrate the weak dependence of the coefficient $c$ on
$\eta_\Vert$, we collapse in Fig.~5 the data obtained for different
$\eta_\Vert$ into one curve by plotting
\begin{equation}
\eta_{ph} = \bar\eta - \eta_\Vert\ .
\end{equation}
The full lines correspond to a quadratic fit of the data points.
According to Figs.~5a and 5b, we deduce for the (100) and (111) faces
\,$c\approx0.6\,t_0^{-1}\epsilon^{-2}\approx 0.42$~ns$^{-1}$meV$^{-2}$, and
\,$c\approx0.8\,t_0^{-1}\epsilon^{-2}\approx 0.56$~ns$^{-1}$meV$^{-2}$,
respectively.
\begin{figure}
\includegraphics[width=7.cm, clip=true]{fig5a_fric.eps}
\includegraphics[width=7.cm, clip=true]{fig5b_fric.eps}
\narrowtext
\caption{Phonon contribution to the sliding friction of
Xe layer on Ag obtained as $\bar\eta - \eta_\Vert$, as a function of
corrugation amplitude $u_0$. Symbols as in Fig.~4. Full
lines are quadratic fits accounting for errors of points. Error bars
are the same as Fig.~4 but are suppressed here for clarity.
(a) (100) surface; (b) (111) surface.}
\end{figure}
The results shown in Figs.~4 and 5 correspond to
\,$\eta_\perp=1/t_0 = 260$~ns$^{-1}$. Simulations for different
values of $\eta_\perp$ in the range
\,$26$~ns$^{-1}\leq\eta_\perp\leq 260$~ns$^{-1}$\, yield nearly
the same overall sliding friction, confirming the results of
previous studies which suggested that the friction in the
direction of the external force is essentially independent of
the thermostat or the single-site friction in the orthogonal
direction~\cite{cieplak94,cieplak94b,PN96}.
The simulations discussed so far were carried out for 68 Xe atoms
within substrate unit cells corresponding to \,$12a\times12a$\, for
Ag(100) and \,$12a\times14a\sqrt3/2$\, for Ag(111). Remarkably, we
observed pronounced finite-size effects associated with the small
dimension of these unit cells. This is illustrated for Ag(100)
in Fig.~6, where $\eta_{ph}$ is plotted as a function of $u_0$.
In general the Xe atoms form approximate hexagonal
layers. Their orientation, however, tends to prefer two angles:
with rows of Xe atoms nearly parallel to the $x$-axis, or nearly
parallel to the $y$-axis. We denote these phases by $\alpha$ and $\beta$,
respectively. Since these orientations persist even in the absence
of the substrate corrugation ($u_0=0$) and for vanishing external
force, it is clear that they are a consequence of the limited size
of the unit cell. The reason is that it is not possible to accomodate
truly incommensurate hexagonal overlayers, at arbitrary angular
orientations, in a rectangular unit cell that is periodically repeated.
\begin{figure}
\includegraphics[width=7.cm, clip=true]{fig6_fric.eps}
\narrowtext
\caption{Phonon contribution to the sliding friction of
Xe layer on Ag(100) obtained as $\bar\eta - \eta_\Vert$, as a function of
corrugation amplitude $u_0$. Symbols and quadratic fits as in Fig.~5.
Full symbols and full line: $\alpha$-phase, open symbols and dashed
line: $\beta$-phase, corresponding to different angular orientation
of Xe monolayer.}
\end{figure}
In order to reduce these finite-size effects, we have performed
simulations for much larger substrate cells (e.g., 288 Xe atoms
in a \, $24a\times24a$\, cell on Ag(100)). Qualitatively, the
net sliding friction for these systems agrees rather well with
the earlier results, but the tendency for orientational alignment
of the Xe layer is greatly reduced.
\subsection{Coverage dependence of sliding friction}
As mentioned above, the coefficient $c$ of the quadratic term in
Eq.~(\ref{c}) depends strongly on overlayer coverage. We illustrate
this point in Fig.~7, where the net sliding friction of the uncompressed
monolayer ($\Theta=0.94$) is compared with that for a fully compressed
Xe layer ($\Theta=1.0$) and for a less dense layer ($\Theta=0.85$).
The compression is seen to lead to a greatly reduced phonon-induced
contribution to $\bar\eta$. On Ag(100)
as well as Ag(111), the magnitude of $c$ is reduced by more than a
factor of two. A strong reduction of dissipation via phonons upon
compression is to be expected since the Xe atoms are much less free
to vibrate about the equilibrium positions.
The opposite trend is found if the coverage is reduced below that of
the uncompressed monolayer. For Xe on Ag(100) at coverage $\Theta=0.85$,
the coefficient $c$ is about \,$1.25$~ns$^{-1}$meV$^{-2}$, i.e., three
times larger than for $\Theta=0.94$. An even more dramatic increase
of $c$ is found on Ag(111). The strong enhancement of $\bar\eta$ is
caused by the higher probability of exciting atomic vibrations in
the less dense overlayer phase.
\begin{figure}
\includegraphics[width=7.cm, clip=true]{fig7a_fric.eps}
\includegraphics[width=7.cm, clip=true]{fig7b_fric.eps}
\narrowtext
\caption{Net sliding friction $\bar\eta$ of Xe layers
on Ag as a function of corrugation amplitude $u_0$ for coverages
$\Theta=0.85, \ 0.94,\ 1.0$. (a) (100) surface; (b) (111) surface.
$\eta_\Vert = 0.65$~ns$^{-1}$.}
\end{figure}
Concerning the overall variation of $\eta_{ph}$ in the coverage
range $\Theta=0.85\ldots1.0$, on Ag(100) for $u_0=0.95$~meV the
variation is about 1~ns$^{-1}$, while for $u_0=0.5$~meV this
variation diminishes to less than $0.3$~ns$^{-1}$, and for
$u_0=1.4$~meV it is larger than 2~ns$^{-1}$. On Ag(111), in the
same coverage range, $u_0=0.3$~meV yields a variation of
$\eta_{ph}$ of about 1~ns$^{-1}$, whereas $u_0=0.2$~meV
($u_0=0.4$~meV) lead to much smaller (larger) variations.
Therefore $\eta_{ph}$ is sensitive to the substrate topology.
In view of the great sensitivity of $\eta_{ph}(\Theta)$ with
respect to the corrugation amplitude, it would be of interest to
have more accurate independent experimental and/or theoretical
determinations of the substrate/adsorbate potential. In
principle, the corrugation amplitude may be estimated from the
lateral vibrational frequency $\omega_\Vert$ of adsorbed Xe,
which can be detected using inelastic He scattering. For Xe on
Cu(111), one finds \,$\omega_\Vert\approx
3\pm1$~cm$^{-1}$~\cite{braun}. Using the relation
\,$u_0=m\omega_\Vert^2/k^2$, one obtains values of $u_0$ in the
approximate range between 0.5 and 2~meV. From Eq.~(\ref{c}) it
follows that $\eta_{ph}$ varies with the fourth power of
$\omega_\Vert$. This is born out by our simulations which show
that such barriers yield a phonon-induced friction in a very
wide range, \,$\eta_{ph}\approx 0.1\ldots5.0$~ns$^{-1}$, if the
coverage is varied between 1.0 and 0.85.
\subsection{Comparison with Experiment}
As pointed out in Section I, the quartz-crystal microbalance
measurements by Krim {\it et al.}~\cite{krim1} for thin Xe films on
Ag(111) show that the slip time \,$\tau=1/\bar\eta$\, exhibits a
characteristic dependence on Xe coverage (see Fig.~1).
Close to one monolayer, the slip time reaches a minimum of about
\,$\tau\approx 0.8\,$~ns and then increases to about 2.0~ns within a
coverage range of less than 0.3~monolayers. In this section we focus on
this coverage range (roughly the ``compression region'') since it
provides the most challenging aspect of the data.
The interpretation of the slip time at submonolayer
coverages in terms of simulations based on small repeated cells is
less reliable because they do not account for the formation of islands.
Also, the presence of defects and steps which are not included in
the theory then plays a larger role. On the other hand, beyond full
monolayer coverage, the observed slip time shows a much weaker coverage
dependence related to additional phonon dissipation in the partly
filled second monolayer.
In Ref. \cite{TS97} it is stated that
the minimum of $\tau$ corresponds to the uncompressed monolayer at
\,$n_a\approx0.0563$~\AA$^{-2}$ and that the subsequent steep rise
reflects the suppression of intra-adsorbate phonon excitations as the
Xe layer becomes more compressed (as noted above, the compressed
phase is only 6~\% denser than the uncompressed layer~\cite{dai94}.)
This coverage assignment implies that the slip time increases
(i)~when the coverage is reduced below that of the uncompressed
monolayer; and (ii)~when the coverage is increased beyond that of the
compressed layer. However, both of these implications are
implausible since in general incommensurate solid layers slide more
easily than fluids~\cite{cieplak94,cieplak94b,PN96}. Thus, the slip time should
decrease rather than increase (i)~when the coverage is lowered below that
of the uncompressed layer and (ii)~when it is increased beyond that of the
compressed layer, i.e., when the second layer is beginning to adsorb.
Since absolute coverage calibration in the experiment is difficult,
there exists a clearly appreciable uncertainty concerning the measured
sliding friction of a monolayer. In fact, it would appear more
plausible to associate the minimum of the slip time not with the
uncompressed layer but with a slightly lower coverage. Part of the
observed steep rise of $\tau$ would then correspond to the reduction
of intra-overlayer phonon processes upon reaching the uncompressed
layer coverage, and the remaining rise should reflect the
additional phonon reduction due to compression. The maximum of the observed slip
time should approximately correspond to the fully compressed phase.
According to these arguments, the uncompressed monolayer might have
$\tau\approx 1.3\ldots1.6$~ns, giving \,$\bar\eta\approx0.6\ldots
0.8$~ns$^{-1}$.
In Fig.~8 the phonon friction $\eta_{ph}(\Theta)$ is plotted as
a function of coverage, in the range
$\Theta=0.85\ldots1.0$,\ for several values of the corrugation
amplitude $u_0$. For ease of comparison with the experimental
results, we include in Fig.~8(b) the measured $\bar\eta=1/\tau$
as a function of coverage in the region near one monolayer where
$\tau$ exhibits the characteristic positive slope. To account
for the uncertainty of the measured coverage, and for the
expected behavior of the sliding friction as discussed above, a
shift of the data to lower coverages, by about
$0.1\ldots0.15$~monolayer, should be allowed for.
\begin{figure}
\includegraphics[width=7.cm, clip=true]{fig8a_fric.eps}
\includegraphics[width=7.cm, clip=true]{fig8b_fric.eps}
\narrowtext
\caption{Calculated phonon friction $\eta_{ph}(\Theta)$
of Xe on Ag as a function of coverage $\Theta$ for several corrugation
amplitudes $u_0$. (a) (100) surface; (b) (111) surface. The dashed lines
are guides to the eye. The error bars indicate the uncertainty due to
fluctuations. Panel (b) also shows the measured sliding friction
$\bar\eta=1/\tau$ in the compression region where the slip time $\tau$
exhibits the characteristic positive slope (see Fig.~1).}
\end{figure}
The coverage
dependence of $\bar\eta$ in the range shown can be characterized
by two important features: the overall magnitude and the slope.
The comparison of $\bar\eta$ with the calculated results shows
that none of the functions $\eta_{ph}(\Theta)$ obtained for
different corrugation amplitudes $u_0$ provides an adequate
description of the data. However, for $u_0=0.19$~meV, the
calculated $\eta_{ph}$ has roughly the right slope even though
its magnitude at full monolayer coverage is far too small. If
$u_0$ is increased to $0.3\ldots0.4$~meV the slope of
$\eta_{ph}(\Theta)$ becomes rapidly larger than that of the
measured $\bar\eta$ while the magnitude near monolayer coverage
is still too small. Only when $u_0$ is increased to much larger
values does $\eta_{ph}$, in the compressed phase, match the observed
$\bar\eta$, but the slope of $\eta_{ph}(\Theta)$ then is much
larger than observed.
As shown in Fig.~8(a), a similar picture is found on Ag(100). Although
the detailed dependence of $\eta_{ph}(\Theta)$ on the corrugation differs
appreciably from the one obtained for the (111) face, near monolayer
coverage, \,$\eta_{ph}$ remains less than $0.1$~ns$^{-1}$ as long as
\,$u_0<1$~meV, with a moderate slope for $\Theta<1$. Only if $u_0$
is increased to \,2~meV we find \,$\eta_{ph}\approx 0.5$~ns$^{-1}$
in the compressed phase. The slope of $\eta_{ph}(\Theta)$, however,
then quickly becomes extremely large.
Thus, our results imply that it is not possible to simultaneously
describe the magnitude {\it and} slope of the observed variation
of $\bar\eta(\Theta)$ exclusively in terms of phonon dissipation.
Only by adding an electronic friction contribution (which in the
narrow range shown in Fig.~8 can be assumed to be independent of
coverage) does the sum \,$\eta_{el}+\eta_{ph}(\Theta)$\, reproduce
the measured variation of $\bar\eta(\Theta)$. The comparison with
the data suggests that \,$\eta_{el}\approx0.5$~ns$^{-1}$ together
with \,$u_0\approx0.2$~meV (total barrier height $\approx0.9$~meV)
provides a qualitative description of the sliding friction measured
on Ag(111) ---if one allows a shift of the data of 0.14~monolayer
to lower coverages, of course. As discussed in Section
II, an overall single-atom electronic friction coefficient of this
magnitude is in agreement with surface resistivity measurements and
independent theoretical estimates.
Interestingly, our findings for Xe/Ag differ from those on
Kr/Au. As recently shown by Robbins and Krim \cite{RK98}, the
Kr/Au QCM data up to monolayer coverage can be fitted rather
well assuming negligible electronic friction. A finite
electronic friction would require a correspondingly smaller
substrate corrugation $u_0$ which, in turn, would yield too low
values of the phonon friction at submonolayer coverages. We
note, however, that the analysis of the low-coverage data is
uncertain since the simulations are limited to repeated finite
cells and do not include the effect of defects or steps, nor do
they properly describe the formation of islands which most
likely occur under experimental conditions. Thus, if one
focusses on the slip time in the less problematic region near
monolayer coverage, i.e., roughly \,$0.85\leq\Theta\leq 1$, the
analysis in Ref.~\cite{RK98} indicates that the inclusion of a
finite electronic friction and a slightly reduced corrugation
$u_0$ is not incompatible with the data.
In addition, we point out that electronic friction of Kr on Au indeed
ought to be weaker than for Xe on Ag for the following reasons. Since
the Au $d$ bands lie only 2~eV below the Fermi level, in contrast to 4~eV
for Ag, the larger $s-d$ hybridization diminishes the polarizability of
the $s$ electron tails decaying into the vacuum region and thereby
reduces the probability of exciting substrate electron--hole pairs
via adsorbed particles. The larger work function of Au (5.3~eV compared
to 4.8~eV) also contributes to a stiffer, less polarizable electronic
density tail \cite{liebsch97,liebsch97b}. Moreover, since the atomic
polarizability of Kr is about 40\,\% smaller than that of Xe \cite{ZS},
the leading van der Waals part of the friction coefficient should
be much smaller than for Xe on Ag. For the same physical reasons,
the contributions associated with the Pauli repulsion and the
covalency of the adatom bond should also be smaller. The overall
electronic coupling of Kr on Au should therefore be significantly
weaker than for Xe on Ag. Thus, great caution is required when
comparing electronic excitation processes on different adsorption
systems ---even in the case of rare gas atoms on noble metal surfaces.
The comparison of Figs.~8(a) and (b) suggests that, for the same
corrugation amplitude $u_0$, Ag(111) yields a larger phonon friction
than Ag(100). This trend seems at first surprising since the
close-packed (111) face is generally regarded as the smoothest of
the low-index faces of an fcc crystal. On the other hand, as pointed
out in Section III, the potential barrier between neighboring hollow
sites is $2u_0$ for the (100) face and $0.5u_0$ for the (111) face,
while the total barrier between hollow sites across the substrate
lattice sites is $4u_0$ for the (100) face and $4.5u_0$ for (111).
Thus, the (111) surface has sharper potential maxima than the (100)
face. Since atoms of an incommensurate overlayer sample not only
the potential minima but the entire substrate surface, it is indeed
plausible that Ag(111) causes more Xe intra layer phonon friction than
Ag(100). According to this, the topology of the substrate is a relevant
feature if one want to make a comparison with the experiments.
In their analysis of an uncompressed Xe layer on Ag(100), Persson
and Nitzan~\cite{PN96} used $u_0=0.95$~meV and obtained
$\eta_{ph}\approx0.01$~ns$^{-1}$ (for \,$\eta_\Vert=0.62$~ns$^{-1}$,
they find $\eta_\Vert/\bar\eta =0.98\pm0.04$). From this result, they
concluded that the net friction at monolayer coverage is mainly of
electronic origin. Instead, for the same parameters, we find
\,$\eta_{ph}\approx 0.3$~ns$^{-1}$, suggesting a phonon friction
roughly half as large as the electronic contribution.
On the other hand, Tomassone {\it et al.}\cite{TS97} assumed a total
barrier height of 2~meV on Ag(111), implying in our notation
\,$u_0=0.45$~meV. On the basis of the simulations it was concluded
that phonon friction dominates over possible electronic contributions.
For an uncompressed Xe layer, these authors obtained a phonon friction
\,$\eta_{ph}\approx1.8$~ns$^{-1}$, which is considerably larger than
what we find for the same parameters. According to Fig.~2 of Ref.
\cite{TS97}, in the compression region, the calculated slip time for
\,$u_0=0.45$~meV increases extremely rapidly in agreement with our
results. This increase, however, is much steeper than what is seen
in the experiment. Thus, a weaker corrugation, combined with a finite
electronic friction might, in fact, provide a better fit of the data
in this coverage range. The analysis of the submonolayer data is
again more complicated for the reasons discussed earlier.
\section{Conclusion}
We have performed molecular dynamics simulations for Xe layers on
Ag(100) and Ag(111) in order to determine the net sliding friction
in the presence of a constant lateral external force. We find that
$\bar\eta$ depends sensitively on the corrugation amplitude $u_0$
of the Xe/Ag interaction potential and on the single-adatom
microscopic friction parameter $\eta_\Vert$. For coverages approaching
one monolayer, the phonon-induced contribution $\eta_{ph}(\Theta)$
decreases rapidly, in agreement with the strong positive slope of the
observed slip time $\tau$. A detailed study of $\eta_{ph}(\Theta)$
as a function of corrugation amplitude $u_0$ suggests that, close to
monolayer coverage, it is not possible to consistently describe both
the magnitude {\it and} slope of the measured $\bar\eta=1/\tau$
without including a constant lateral electronic friction $\eta_{el}$.
Our simulations indicate that this electronic contribution should be
about 0.5 ns$^{-1}$ which is in qualitative agreement
with surface resistivity measurements and independent theoretical
estimates.
The overall picture that
seems to be consistent with the available information is that the
rapidly varying part of the observed sliding friction is due to
excitation of atomic vibrations within the overlayer and that
single-site electronic friction processes contribute a roughly
coverage-independent background. For a fully compressed Xe layer,
phonon dissipation is weaker than the electronic friction. As
the coverage is reduced, however, phonon processes rapidly become
important and eventually, near the minimum of the observed slip
time, dominate over the electronic dissipation channel.
In summary, even though both experimental and theoretical
results on sliding friction of Xe on Ag have inherent
uncertainties, our analysis indicates that both mechanisms
---phononic and electronic dissipation--- do contribute. The
relative weight of these channels depends strongly on coverage
in the crucial region near a full monolayer.
\acknowledgments
One of the authors (A.L.) gratefully acknowledges
the hospitality of the Facultad de F\'\i sica, Universidad
Cat\'olica, Chile, where this work was partially carried out, as well
as financial support by the European Community. A. L. also thanks
Prof. Abe Nitzan for making available his molecular dynamics code and
Dr. P. Ballone for several useful suggestions concerning the simulations.
M. K. was partially supported by the {\it Fondo Nacional de
Investigaciones Cient\'\i ficas y Tecnol\'ogicas} (FONDECYT, Chile)
under grant \#~1971212 and S. G. under grant \#~3950028.
\noindent
$^*$e-mails:
<EMAIL>, \ <EMAIL>, \ <EMAIL>
|
\section{Introduction}
\subsection{The standard formalism}
The standard formalism for the
calculation of bubble-nucleation rates during first-order
phase transitions in field theories at non-zero temperature
was introduced in refs.~\cite{coleman}--\cite{linde}.
It consists of an implementation of
Langer's theory of homogeneous
nucleation~\cite{langer} (for a review see ref.~\cite{review})
within the field-theoretical context.
The nucleation rate $I$
gives the probability per unit time and volume to nucleate a certain
region of the stable phase (the true vacuum) within the metastable
phase (the false vacuum).
Its calculation relies on a semiclassical approximation
around a dominant saddle-point, which is identified with
the critical bubble.
This is a static configuration
(usually assumed to be spherically symmetric) within the metastable phase
whose interior consists of the stable phase.
It has a certain radius that can be determined from the
parameters of the underlying theory. Bubbles slightly larger
than the critical one expand rapidly, thus converting the
metastable phase into the stable one.
The nucleation rate is exponentially suppressed by the action
(the free energy rescaled by the temperature) of the critical bubble.
Possible deformations of the critical bubble
generate a pre-exponential factor.
The leading contribution to this factor
has the form of a ratio of fluctuation determinants and corresponds to the
first-order correction to the semiclassical result in a systematic
expansion around the saddle point.
For a $(d+1)$-dimensional theory of a real scalar field
at temperature
$T$, in the limit that thermal fluctuations dominate over
quantum fluctuations, the bubble-nucleation rate
is given by~\cite{colcal}--\cite{linde}
\begin{equation}
I=\frac{E_0}{2\pi}
\left(\frac{S}{2\pi }\right)^{d/2}\left|
\frac{\det'[\delta^2 \Gamma/\delta\phi^2]_{\phi=\phi_{\rm b}}}
{\det[\delta^2 \Gamma/\delta\phi^2]_{\phi=0}}\right|^{-1/2}
\exp\left(-S\right).
\label{rate0} \end{equation}
Here $\Gamma$ is the free energy of the system for a given configuration of the
field $\phi$.
The rescaled
free energy of the critical bubble is $S=\Gamma_b/T
=\left[\Gamma\left(\phi_{\rm b}(r)\right)-\Gamma(0)\right]/T$,
where $\phi_{\rm b}(r)$ is the spherically-symmetric
bubble configuration and $\phi = 0$ corresponds to the false vacuum.
The prime in the fluctuation determinant around
the bubble denotes that the $d$ zero eigenvalues
of the operator $[\delta^2 \Gamma/\delta\phi^2]_{\phi=\phi_{\rm b}}$,
corresponding to displacements of the bubble,
have been removed.
Their contribution generates the factor
$\left(S/2\pi \right)^{d/2}$ and the volume factor
that is absorbed in the definition of $I$ (nucleation rate per unit volume).
The quantity $E_0$ is the square root of
the absolute value of the unique negative eigenvalue.
In field theory, the free energy
density $\Gamma$ of a system for homogeneous configurations
is usually identified with
the temperature-dependent effective potential. This is evaluated
through some perturbative scheme, such as the loop expansion.
The profile and the free energy of the critical bubble are determined
through the potential. This approach, however,
faces fundamental difficulties:
For example, the effective potential, being the Legendre transform of the
generating functional for the connected Green functions,
is a convex function of the field. Consequently, it does not
seem to be the appropriate quantity for the study of tunnelling.
Also,
the fluctuation determinants in the expression for the nucleation
rate have a form completely analogous to the one-loop correction to
the potential. The question of double-counting the effect of
fluctuations (in the potential and the prefactor)
must be properly addressed.
A closely related issue
concerns the ultraviolet divergences that are inherent in
the calculation of the fluctuation determinants
in the prefactor. An appropriate regularization scheme must be
employed in order to control them~\cite{previous}.
Moreover, this scheme must be consistent with the one employed
for the absorption of the divergences appearing in the
calculation of the potential.
\subsection{Coarse-graining}
In refs.~\cite{bubble}--\cite{third}
it was shown that all the above issues can be resolved
through the implemention of the notion of coarse graining in the
formalism. The appropriate quantity for the description of the
physical system is
the effective average action $\Gamma_k$~\cite{averact}, which
is the generalization in the continuum
of the blockspin action
of Kadanoff~\cite{kadanoff}. It can be interpreted as a coarse-grained
free energy at a given scale $k$.
Fluctuations with
characteristic momenta $q^2 \gta k^2$ are integrated out
and their effect is incorporated in
$\Gamma_k$.
In the limit $k \to 0$,
$\Gamma_k$ becomes equal to the effective action.
The $k$ dependence
of $\Gamma_k$ is described by an exact flow equation~\cite{exact},
typical of the Wilson approach to the renormalization group~\cite{wilson}.
This flow equation can be translated into evolution equations
for the functions appearing in a derivative expansion of
the action~\cite{indices,morris}.
Usually, one
considers only the effective average potential
$U_k$ and a standard kinetic term, and neglects higher derivative
terms in the action. We shall employ this approximation
in this paper also.
The bare theory is defined
at some high scale $\Lambda$ that can be identified with the ultraviolet
cutoff.
At scales $k$ below the temperature $T$,
a $(d+1)$-dimensional theory at non-zero temperature
can be described in terms of an effective
$d$-dimensional action at zero temperature~\cite{trans}.
In ref.~\cite{first} we considered a (3+1)-dimensional
theory of a real scalar field at non-zero temperature,
defined through its action
$\Gamma_{k_0}$
at a scale $k_0$ below the temperature, so that
the theory has an effective three-dimensional description.
The form of the potential $U_{k_0}$
results from the bare potential
$U_{\Lambda}$ after the integration of (quantum and thermal)
fluctuations between the scales
$\Lambda$ and $k_0$.
We computed the form of the $U_k$ at scales $k\leq k_0$ by
integrating an evolution equation derived from the exact flow
equation for $\Gamma_k$.
$U_k$ is non-convex for non-zero $k$, and
approaches convexity only in the limit $k\to 0$.
The nucleation rate must be computed for $k$ larger than the scale $k_f$
at which the functional integral in the definition of $U_k$ starts
receiving contributions from field configurations that interpolate between
the two minima. This happens when $-k^2$ becomes approximately equal to
the negative curvature at the top
of the barrier~\cite{convex}.
For $k \gta k_f$ the typical length scale of a thick-wall critical
bubble is $\gta 1/k$.
We performed the calculation of the nucleation rate for a range of scales
above and near $k_f$, for which $U_k$ is non-convex.
In our approach
the pre-exponential factor is well-defined and finite,
as an ultraviolet cutoff of order $k$ is implemented in the calculation of
the fluctuation determinants, so that fluctuations
with characteristic momenta $q^2 \gta k^2$ are not included.
This is a natural consequence of the fact
that all fluctuations with typical momenta above $k$ are
already incorporated in the form of $U_k$.
This modification also resolves naturally the problem of double-counting
the effect of the fluctuations.
We found that the saddle-point configuration
has an action $S_k$ with a significant $k$ dependence.
For strongly first-order phase transitions, the nucleation
rate $I = A_k \exp(-S_k)$
is dominated by the exponential suppression.
The main role of the prefactor $A_k$, which is also $k$ dependent,
is to remove the scale dependence from the total nucleation rate.
Thus, this physical quantity is independent of the scale $k$ that we
introduced as a calculational tool.
The implication of our results
is that the critical bubble should not be identified
just with the saddle point of the semiclassical approximation.
It is the combination of
the saddle point and its possible deformations
in the thermal bath (accounted for by the fluctuation
determinant in the prefactor) that has physical meaning.
We also found that,
for progressively more weakly first-order phase transitions,
the difference between
$S_k$ and $\ln ( A_k/k^4_f )$ diminishes.
This indicates that the effects of fluctuations
become more enhanced.
At the same time, a significant $k$ dependence of the
predicted nucleation rate develops.
The reason for the above deficiency is clear.
When the nucleation rate is roughly equal to or smaller
than the contribution from the prefactor, the effect of
the next order in the expansion around the saddle point is
important and can no longer be neglected.
This indicates that there is a limit to the validity of Langer's picture
of homogeneous nucleation. The region of validity of this picture
was investigated in detail in ref.~\cite{second}. First-order phase transitions
in two-scalar models were studied in ref.~\cite{third} and the consistency of
the approach summarized above was reconfirmed.
Moreover, the applicability of homogeneous nucleation theory to
radiatively-induced first-order phase transitions was tested.
It was found that the expansion around the semiclassical saddle point
is not convergent for such phase transitions. This indicates that
estimates of bubble-nucleation rates for the electroweak phase transition
that are based only on the saddle-point action may be very misleading.
\subsection{Plan of the paper}
In this paper we present an application of our formalism to
(2+1)-dimensional theories at non-zero temperature. Our investigation provides
a new test of several points of our approach that depend strongly on the
dimensionality, such as the form of the
evolution equation of the potential, the nature of the
ultraviolet divergences of the fluctuation determinants,
and the $k$ dependence of
the saddle-point action and prefactor. The
complementarity between the $k$ dependence of $S_k$ and $A_k$ is a
crucial requirement for the nucleation rate $I$ to be $k$ independent.
A strong motivation for this study stems from the existence of
lattice simulations of nucleation for (1+1) and (2+1)-dimensional systems
\cite{onepone,twopone}. In particular, we shall compare our predictions
for the nucleation rate with the lattice results of ref.~\cite{twopone}.
In the next section we summarize the basic steps of our method
and derive the necessary expressions for the calculation of the
nucleation rate. In section 3 we present sample calculations
in two dimensions. In section 4 we apply our formalism to theories
that have been studied through lattice simulations and compare with
the lattice results. Our conclusions are presented in section 5.
\setcounter{equation}{0}
\renewcommand{\theequation}{\thesection.\arabic{equation}}
\section{The calculation of bubble-nucleation rates}
\subsection{Evolution equation for the potential}
We consider a model of a real scalar field $\phi$ in 2+1 dimensions.
The
effective average action $\Gamma_k(\phi)$~\cite{averact}
is obtained by
adding an infrared cutoff term to the bare action,
so that contributions from modes with
characteristic momenta $q^2 \lta k^2$ are not taken into account.
We use
the simplest choice of a mass-like cutoff term $\sim k^2 \phi^2$,
for which the
perturbative inverse propagator for massless fields is
$P_k(q) \sim q^2 + k^2$. This choice makes
the calculation of the fluctuation determinants in the pre-exponential factor
of the nucleation rate technically feasible.
Subsequently,
the generating functional for the connected Green functions is defined,
from which the
generating functional for the 1PI Green functions can be obtained through
a Legendre transformation.
The presence of the modified propagator in the above definitions
results in the effective integration of the
fluctuations with $q^2 \gta k^2$ only. Finally, the
effective average action is
obtained by removing the infrared cutoff from the
generating functional for the 1PI Green functions.
The effective average action $\Gamma_k$ obeys an exact
flow equation,
which describes its response to variations of the infrared cutoff $k$
~\cite{exact}. This can be turned into evolution equations for the
functions appearing in a derivative expansion of $\Gamma_k$~\cite{indices}.
In this work we use an approximation which neglects higher derivative
terms in the action and approximates it by
\begin{equation}
\Gamma_k =
\int d^2x \left\{
\frac{1}{2} \partial^{\mu} \phi~\partial_{\mu} \phi~
+ U_k(\phi) \right\}.
\label{twoeleven} \end{equation}
The above action describes the effective two-dimensional theory that
results from the dimensional reduction of a high-temperature
(2+1)-dimensional theory at scales below the temperature.
The temperature has been absorbed in a redefinition of the fields and
their potential, so that these have dimensions appropriate
for an effective two-dimensional theory. The correspondence
between the quantities we use and the ones of the (2+1)-dimensional
theory is given by
\begin{eqnarray}
\phi =\frac{\phi_{2+1}}{\sqrt{T}},\qquad
U(\phi)=\frac{U_{2+1}\left( \phi_{2+1},T \right)}{T}.
\label{fivethree} \end{eqnarray}
In this way, the temperature does not appear explicitly in
our expressions. This has the additional advantage of
permitting the straightforward application of our results
to the problem of quantum tunnelling in a two-dimensional
theory at zero temperature~\cite{first}.
The evolution equation for the potential
can be written in the form~\cite{exact,indices,second}
\begin{equation}
\frac{\partial}{\partial k^2} \left[ U_k(\phi) - U_k(0)\right] =
-\frac{1}{8 \pi} \left[
\ln \left( 1 + \frac{U''_k(\phi)}{k^2} \right)
- \ln \left( 1 + \frac{U''_k(0)}{k^2} \right) \right].
\label{evpot} \end{equation}
For the numerical integration of the above equation we use the
algorithms described in ref.~\cite{num}.
The first step of an iterative solution of eq.~(\ref{evpot}) gives
\cite{iterative}
\begin{equation}
U_k^{(1)}(\phi)-U_k^{(1)}(0)=
U_{k_0}(\phi)-U_{k_0}(0)
+
\frac{1}{2}\ln\left[\frac{\det[-\partial^2+k^2+U''_{k_1}(\phi)]}{
\det[-\partial^2+k^2_0+U''_{k_1}(\phi)]}
\frac{\det[-\partial^2+k^2_0+U''_{k_1}(0)]}{
\det[-\partial^2+k^2+U''_{k_1}(0)]}\right].
\label{iter} \end{equation}
The scale $k_1$ in the above expression can be chosen
arbitrarily anywhere between $k_0$ and $k$, as the induced
uncertainty for $U_k^{(1)}(\phi)$ corresponds to a higher-order
contribution in the iterative procedure.
For $k_1=k_0$, $k\to 0$, eq.~(\ref{iter}) is a regularized one-loop
approximation to the effective
potential. Due to the ratio of determinants, only
momentum modes with $k^2\lta q^2 \lta k_0^2$ are effectively included
in the momentum integrals.
The above expression demonstrates the form of ultraviolet regularization
of fluctuation determinants that is consistent with the cutoff
procedure that leads to the evolution equation for the potential. An analogous
regularization must be used for the fluctuation
determinants in the expression for the nucleation rate.
\subsection{The nucleation rate}
The calculation of the nucleation rate proceeds in complete analogy
to the one in ref.~\cite{first} which we outlined in the introduction.
We define the theory through the potential at a scale $k_0$ below the
temperature, so that the behaviour is effectively two-dimensional.
We then integrate the evolution equation down to a scale
$k \gta k_f$, where we compute the bubble-nucleation rate.
In practice, $k_f^2$ is taken $10\%$ larger than the absolute value of
the curvature at the top of the barrier.
The potential has two minima:
the stable (true)
one located at $\phi=\phi_t$, and the unstable (false) one at
$\phi=\phi_f=0$.
The nucleation rate is exponentially suppressed by
the action $S_k$ (the free energy rescaled by the temperature)
of the saddle-point configuration
$\phi_{\rm b}(r)$ that is associated with tunnelling.
This is an $SO(2)$-symmetric
solution of
the classical equations of motion
which interpolates between
the local maxima of the potential
$-U_k(\phi)$. It satisfies the equation
\begin{equation}
{d^2\phi_{\rm b}\over dr^2}+\frac{1}{r}~{d\phi_{\rm b}\over dr}=
U'_k(\phi_{\rm b}),
\label{eom} \end{equation}
with the boundary conditions
$\phi_{\rm b}\rightarrow 0$ for $r \rightarrow \infty$ and
$d\phi_{\rm b}/dr= 0$ for $r =0$.
The action $S_k$ of the saddle point is given by
\begin{equation}
S_k=2 \pi
\int_0^\infty
\left[ \frac{1}{2}
\left(\frac{d\phi_{\rm b}(r)}{dr}\right)^2
+U_k(\phi_{\rm b}(r)) -U_k(0) \right]
r\,dr\equiv S_k^t+S_k^v.
\label{action} \end{equation}
The profile of the saddle point can be easily computed
with the ``shooting'' method~\cite{shooting}.
A consistency check for our solution is provided by the fact
that $S_k^v=0$ for two-dimensional theories.
The bubble-nucleation rate is determined through eq.~(\ref{rate0})
in terms of the potential $U_k(\phi)$. The explicit temperature dependence
is absorbed in the definition of effective two-dimensional parameters
according to eq.~(\ref{fivethree}).
An appropriate regularization is implemented in order to control the
ultraviolet divergence of the prefactor.
The type of regularization
is dictated by the one-loop effective potential,
given by eq.~(\ref{iter}) in our scheme.
This equation indicates that fluctuation
determinants computed within the low-energy theory must be
replaced by appropriate ratios of determinants.
We emphasize that the matching of the regularization scheme for the
prefactor with the cutoff used for the derivation of the
evolution equation for the potential~(\ref{evpot}) is crucial for the
consistency of our method.
Finally, the nucleation rate in two dimensions is given by
\begin{eqnarray}
I&=&A_{k} \exp({-S_k})
\nonumber \\
\riga{where}\\[-2mm]
A_{k}&=& \frac{E_0}{2\pi} \frac{S_k}{2\pi}
\left|
\frac{\det'\left[-\partial^2+U''_k(\phi_{\rm b}(r)) \right]}
{\det \left[ -\partial^2+k^2 +U''_k(\phi_{\rm b}(r))\right]}
~\frac{\det\left[-\partial^2+k^2+U''_k(0) \right]}
{\det\left[-\partial^2+U''_k(0)\right]}
\right|^{-1/2}.
\label{rrate} \end{eqnarray}
The differential operators that appear in eq.~(\ref{rrate})
have the general form
\begin{eqnsystem}{sys:W}
{\cal W}_{\kappa\alpha}&=&-\partial^2 +m_{\kappa}^2+\alpha W_{k}(r)
\label{op} \\
\riga{where}\\[-2mm]
m_{\kappa}^2&\equiv&U''_k(0)+\kappa k^2,\\
W_{k}(r)&\equiv&U''_k(\phi_{\rm b}(r))-U''_k(0),\label{ak}
\end{eqnsystem}
with $\kappa,\alpha=0$ or 1.
As the ${\cal W}_{\kappa\alpha}$ operators are $SO(2)$-symmetric,
it is convenient to use polar coordinates and express their
eigenfunctions as $\psi_n(r,\varphi)=e^{in\varphi} u_n(r)/\sqrt{r}$.
This leads to
\begin{eqnarray}
\det {\cal W}_{\kappa\alpha}
&=&\prod_{n=-\infty}^\infty \det {\cal W}_{n\kappa\alpha}
\nonumber \\
{\cal W}_{n\kappa\alpha}&=&-\frac{d^2}{dr^2}
+\frac{n^2-\frac{1}{4}}{r^2}+m_{\kappa}^2+\alpha W_{k}(r).
\label{opl} \end{eqnarray}
The computation of such determinants is made possible by
a theorem~\cite{erice} that relates ratios
of determinants to solutions of ordinary differential equations.
In particular, we have
\begin{equation}
g_{n\kappa}\equiv
\frac{\det {\cal W}_{n \kappa 1}}{\det {\cal W}_{n \kappa 0}}=
\frac{y_{n\kappa1}(r\to \infty)}{y_{n\kappa0}(r \to \infty)},
\label{theorem} \end{equation}
where $y_{n\kappa\alpha}(r)$ is the solution of the differential equation
\begin{equation}
\left[-\frac{d^2}{dr^2}
+\frac{n^2-\frac{1}{4}}{r^2}
+m_{\kappa}^2+\alpha W_{k}(r)\right]y_{n\kappa\alpha}(r)=0,
\label{diffeq} \end{equation}
with the behaviour
$y_{n\kappa\alpha}(r)\propto r^{|n|+\frac{1}{2}}$ for $r\to 0$.
For example, $y_{n\kappa0}$ are proportional to modified Bessel functions:
$y_{n\kappa0}\propto \sqrt{r}~I_{|n|}(m_\kappa r).$
Equations such as~(\ref{diffeq}) can be solved numerically with
Mathematica~\cite{Mathematica}.
Since opposite values of $n$ lead to identical determinants,
the final expression for the nucleation rate can be written as
\begin{eqnarray}
I &=& \frac{1}{2 \pi}
\left(\frac{S_k}{2\pi}\right)\exp\left(-S_k\right)
\prod_{n=0}^\infty c_{n},
\nonumber \\
c_{0} &=& \left( \frac{E_0^2 g_{01}}{\left| g_{00} \right|}
\right)^{1/2},~~~~~~~~~~
c_{1} = {g_{11}\over g'_{10}},~~~~~~~~~~
c_{n} = {g_{n 1}\over g_{n 0}}.
\label{fin} \end{eqnarray}
The calculation of $c_{1}$ is slightly complicated because of the necessity to
eliminate the zero eigenvalue in $g'_{10}$.~(The two zero eigenvalues of
the operator $-\partial^2+U''_k(\phi_{\rm b}(r))$ are included in the
equal factors $g_{-10}$ and $g_{10}$).
Also the (unique) negative eigenvalue $-E_0^2$
must be computed for the determination of $c_{0}$.
How these steps are achieved is described in ref.~\cite{first}.
For sufficiently large $n$, one can compute $c_n$ analytically using
first-order perturbation theory in $W_k(r)$ \cite{previous,first}.
We find
\begin{eqnarray}
g_{n\kappa} \approx 1+\frac{1}{2n} \int r\,W_k(r)\,dr,\qquad
c_n \approx 1-\frac{1}{4n^3}k^2\int r^3 \, W_k(r)\,dr.
\label{apppr} \end{eqnarray}
These expressions are very useful for the evaluation of the prefactor, as
only $c_n$ for small values of $n$ need to be computed numerically.
\setcounter{equation}{0}
\renewcommand{\theequation}{\thesection.\arabic{equation}}
\begin{figure}[t]
\begin{center}\hspace{-5mm}
\begin{picture}(17,9)
\putps(0,-0.4)(-2.1,-0.5){figEx}{figEx.ps}\special{color cmyk 0 1. 1. 0.5}
\put(2.5,8.5){\makebox(0,0){{\bf (a)}}}
\put(6.83,8.5){\makebox(0,0){{\bf (b)}}}
\put(11.16,8.5){\makebox(0,0){{\bf (c)}}}
\put(15.5,8.5){\makebox(0,0){{\bf (d)}}}
\special{color cmyk 0 0 0 1.}
\put(2.5,-0.5){\makebox(0,0){$\phi$}}
\put(6.83,-0.5){\makebox(0,0){$r$}}
\put(11.16,-0.5){\makebox(0,0){$r$}}
\put(15.5,-0.5){\makebox(0,0){$\ln k$}}
\end{picture}
\vspace{1cm}
\caption[SP]{\em The steps in the computation of the nucleation rate:
{\rm (a)} Potentials $V_k(\phi)$;
{\rm (b)} Saddle points $\phi_b(r)$;
{\rm (c)} $W_k(r)$, given by eq.~(\ref{ak});
{\rm (d)} Results for the saddle-point action
$S_k$ (diamonds), prefactor $\ln\left(A_k/m^3\right)$ (stars) and nucleation
rate $-\ln \left(I/m^3\right)$ (squares).
The first row corresponds to a model with a potential $U_{k_0}(\phi)$ given
by eq.~(\ref{eq:two20}) with $\gamma/m^2=-2$,
$h/m^2=1/3$, while
the second one to a model with $\gamma/m^2=-4.5$,
$h/m^2=2$.
All dimensionful quantities are given in units of $m$.
\label{fig:Ex}}
\end{center}\end{figure}
\section{A sample computation}
We are interested in potentials that have the approximate form
\begin{equation}
V(\phi) =
\frac{m^2}{2} \phi^2
+\frac{\gamma}{6} \phi^3
+\frac{h}{8} \phi^4.
\label{eq:two20} \end{equation}
The explicit
temperature dependence has been absorbed in $\gamma$,
$h$ and $\phi$ according to eqs.~(\ref{fivethree}).
Through a shift $\phi\to\phi+c$ the cubic term in eq.~(\ref{eq:two20})
can be eliminated
in favour of a term linear in $\phi$.
The resulting potential
describes a statistical system of the Ising universality class
in the presence of an external magnetic field.
We assume that the potential of eq.~(\ref{eq:two20}) describes the
theory at some initial scale $k_0$ not far below the temperature,
similarly to refs.~\cite{first}--\cite{third}. Its form
at lower scales can be determined by integrating the evolution equation
(\ref{evpot}) numerically, or by using the approximate
solution of eq.~(\ref{iter}).
The various steps in our calculation are summarized in fig.~1,
for two models described by a potential $U_{k_0}(\phi)$ given
by eq.~(\ref{eq:two20}).
In the following we express all dimensionful quantities in terms of the
arbitrary mass scale $m$.
In the first row we present results for a theory
with
$\gamma/m^2=-2$,
$h/m^2=1/3$.
In (a)
we present the evolution of the potential $U_k(\phi)$ as the scale
$k$ is lowered.
We always shift the metastable vacuum to
$\phi=0$.
The solid line corresponds to $k_0/m=2$, while the
line with longest dashes (that has the smallest barrier height)
corresponds to $k_f/m=1.1$. At the scale $k_f$ the negative
curvature at the top of the barrier is slightly larger than
$-k_f^2$.
This is the point in the evolution of the potential
where configurations that
interpolate between the minima start becoming relevant
in the functional integral that defines the coarse-grained potential
\cite{convex}.
For this reason,
we stop the evolution at this point.
We observe that, for low $k$, the absolute minimum of the potential
settles at a non-zero value of $\phi$. A significant barrier
separates it from the metastable minimum at $\phi=0$.
The profile of the saddle point $\phi_{\rm b}(r)$
is plotted in (b) in units of $m$
for the same sequence of scales.
We observe a
variation of the value of the field $\phi$ in the center of the critical
bubble for different $k$.
This is reflected in the form of the quantity $W_k(r)$, defined in
eq.~(\ref{ak}), which we plot in~(c).
Our results for the nucleation rate are presented in (d).
On the horizontal axis we give the values of $\ln(k/m)$.
The dark diamonds correspond to the values of the action $S_k$
of the saddle point at the scale $k$.
The stars indicate the values of
$\ln ( A_k/m^3 )$.
We observe a logarithmic $k$ dependence
for both these quantities.
The value of $A_k$ is expected to decrease for smaller $k$, because
$k$ acts as the effective ultraviolet cutoff in the calculation
of the fluctuation determinants in $A_k$. For smaller $k$,
fewer fluctuations with wavelengths above an increasing length scale
$\sim 1/k$ contribute explicitly to the fluctuation determinants.
The logarithmic dependence on $k$ is the reflection of the logarithmic
ultraviolet divergence of the unregularized prefactor in two dimensions.
The dark squares in (d) give our results for
$-\ln(I/m^3)
= S_k-\ln ( A_k/m^3 )$.
The logarithmic $k$ dependence largely cancels between
$S_k$ and $\ln ( A_k/m^3 )$, so that
$\ln(I/m^3)$ is almost constant.
The small residual dependence on $k$ can be used to estimate the
contribution of the next order in the expansion around the saddle point.
This contribution is expected to be smaller than
$\ln ( A_k/m^3 )$.
This behaviour confirms that the
nucleation rate should be independent of the scale $k$ that
we introduced as a calculational tool.
In the second row we present the calculation of the nucleation rate for
a model with a larger coupling $h/m^2=2$, and
$\gamma/m^2=-4.5$,
$k_0/m=2$, $k_f/m=1.1$.
We observe a more pronounced $k$ dependence of the potential, saddle-point
profile and function $W_k(r)$.
The most important aspect of the comparison of the two models
concerns the relative values of
$S_k$ and $\ln ( A_k/m^3 )$.
For the first model,
the contribution of the prefactor to the
nucleation rate is much smaller than that of the action of the
saddle point. The main role of the prefactor is
to remove the logarithmic $k$ dependence from $I/m^3$.
For the second model,
$S_k$ and $\ln ( A_k/m^3 )$ are comparable.
This indicates that the effects of fluctuations
are enhanced.
Moreover, the prefactor fails to cancel the $k$ dependence
of the saddle-point action.
The reason is that
the next order in the expansion around the saddle point is
important and can no longer be neglected.
This establishes the limit of validity of
homogeneous nucleation theory~\cite{langer} for two-dimensional
systems, in agreement with the studies of refs.~\cite{first}--\cite{third}
in three dimensions.
\section{Comparison with lattice studies}
\subsection{Matching the lattice theory}
A main objective of this work is the comparison of our results
with the lattice study of ref.~\cite{twopone}.
This requires a precise definition of the form of the potential.
We must make sure that we consider a theory identical to
the one simulated on the lattice.
As we cannot match exactly the bare parameters of the lattice action, it
is more convenient to guarantee that the low energy renormalized theory
is the same in our model and ref.~\cite{twopone}.
In the latter work,
through a redefinition of the field and the distance, the
one-loop renormalized action (free energy rescaled with respect to the
temperature)
of the (2+1)-dimensional theory at high temperature
is expressed as
\begin{equation}
S=
\frac{1}{\theta} \int d^2 \tilde{x} \biggl\{
- \frac{1}{2} \left( 1-\frac{\theta}{48\pi}\right)
\tilde{\phi} \, \tilde{\partial}^2 \tilde{\phi}
+\tilde{V}(\tilde{\phi}) + \frac{\theta}{8\pi}\left[
\tilde{V}''(\tilde{\phi})-\tilde{V}''(\tilde{\phi})\ln \left(\tilde{V}''(\tilde{\phi})\right)
\right] \biggr\},
\label{renorm} \end{equation}
with
\begin{equation}
\tilde{V}(\tilde{\phi})= \frac{1}{2}\tilde{\phi}^2 - \frac{1}{6}\tilde{\phi}^3+ \frac{1}{24} \lambda \tilde{\phi}^4.
\label{pot} \end{equation}
For $\lambda<0$ the potential becomes unbounded from below, while for
$\lambda=1/3$ it has two equivalent minima. For this reason, we
consider $\lambda$ values in the interval $(0,1/3)$.
Higher-loop corrections are expected to be proportional to larger powers
of $\theta/8\pi$. This implies that
the theory that is simulated on the lattice corresponds to a renormalized
action in the continuum given by eqs.~(\ref{renorm}), (\ref{pot}) only if
$\theta$ is not much larger than 1. In the opposite case, the renormalized
action cannot be determined perturbatively.
In our approach the theory is defined at some initial scale $k_0$ and the
integration of evolution equations such as eq.~(\ref{evpot}) generates
the low-energy structure. For $k=0$ the effective average action of
eq.~(\ref{twoeleven}) must be identified with the action of eq.~(\ref{renorm}).
We neglect wavefunction renormalization effects,
which can be seen from eq.~(\ref{renorm})
to be a good approximation for values of $\theta$ not much larger than 1.
The form of the potential at some non-zero scale
$k$ can be inferred by
demanding that the iterative solution of eq.~(\ref{iter}) reproduces the
potential of eq.~(\ref{renorm}) for $k=0$.
In particular, by choosing $k=k_1=0$ in
eq.~(\ref{iter}) and renaming $k_0$ as $k$ we find
\begin{equation}
U_{k}(\phi) \approx V(\phi) + \frac{1}{8\pi} V''(\phi)
- \frac{1}{8\pi} \left(k^2+V''(\phi) \right)
\ln \left(\frac{k^2+V''(\phi)}{m^2}
\right),
\label{incond} \end{equation}
where $V(\phi)$ is given by eq.~(\ref{eq:two20}) with
\begin{eqnarray}
\frac{\gamma}{m^2}=-\sqrt{\theta},\qquad
\frac{h}{m^2}=\frac{1}{3}\, \theta\, \lambda.
\label{param} \end{eqnarray}
In eq.~(\ref{incond}) we have neglected
terms $\sim(8 \pi)^{-2}$ and
terms $\sim (8 \pi)^{-1}$
in the arguments of the logarithms.
It is clear from the above that the dimensionless coupling that controls the
validity of the perturbative expansion is $\theta \lambda/3$. For this reason,
perturbation theory is expected to break down for $\theta \gta 3 /\lambda$.
In summary, eq.~(\ref{incond}) is an approximate solution of the evolution
equation (\ref{evpot}) (at the first level of an iterative procedure)
which is consistent with
eq.~(\ref{renorm}) that determines the renormalized action of
ref.~\cite{twopone}.
The matching between the renormalized and the lattice actions
is accurate only at the one-loop level
(the region of validity of eq.~(\ref{renorm})).
The approximate
solution of eq.~(\ref{incond}), which results at the first level of the
iterative procedure, has a similar
region of applicability.
Morever, this approximation
is not valid for values of
$k$ that render negative the argument of the logarithm in the
right-hand side of eq.~(\ref{incond}), i.e. for
$k^2 < {\rm max} \left\{-V''(\phi)\right\}$.
This implies that
eq.~(\ref{renorm}) is trustable only in the convex regions of
the potential (for which $\tilde{V}''(\phi) > 0$)
and should not
be expected to lead to reliable predictions for the nucleation rate. This
is in agreement with the conclusion of
ref.~\cite{twopone} that
the determination of the bubble-nucleation rate
through the real part of
eq.~(\ref{renorm}) does not lead to consistency
with the lattice results. On the other hand,
eq.~(\ref{renorm}) is perfectly valid in the convex regions of the
potential, where it can be matched with
eq.~(\ref{incond}) for $k = 0$.
In the following, for our predictions of the bubble-nucleation rates and
the comparison with the lattice results, we
rely on the approximate solution of eq.~(\ref{incond}),
instead of integrating eq.~(\ref{evpot}) numerically.
The numerical solution, which is more accurate than eq.~(\ref{incond}),
does not offer increased
precision in our comparison with the lattice data, as the determination of
the renormalized theory is valid only at the
one-loop level\footnote{We also perform checks of the corrections arising from
integrating the full evolution equation
(\ref{evpot}). These corrections are very small in
the parameter region for which the expansion around the saddle point
is convergent.}.
\subsection{Comparison with the lattice results}
In fig.~2 we present a comparison of results obtained through our method
with the lattice results of fig.~1 of ref.~\cite{twopone}.
For each of several values of $\lambda$ we vary the parameter $\theta$
and determine the couplings $\gamma$, $h$ according to eqs.~(\ref{param}).
The coarse-grained potential is then given by eq.~(\ref{incond}) for
$k \geq k_f$.
The diamonds denote the saddle-point action $S_k$.
For every choice of $\lambda$, $\theta$ we determine $S_k$
at two scales: $1.2\,k_f$ and $2\,k_f$. The light-grey region between
the corresponding points gives an indication of the
$k$ dependence $S_k$.
The bubble-nucleation rate $-\ln\left(I/m^3\right)$ is denoted
by dark squares.
The dark-grey region between the values obtained at
$1.2\,k_f$ and $2\,k_f$ gives a good check of the convergence of the
expansion around the saddle-point. If this region is thin, the
prefactor is in general small and
cancels the $k$ dependence of the action.
The dark circles denote the results for the nucleation rate from
the lattice study of ref.~\cite{twopone}. The dashed straight lines
correspond to the action of
the saddle-point computed from the `tree-level'
potential of eq.~(\ref{eq:two20}).
\begin{figure}[t]
\begin{center}\hspace{-5mm}
\begin{picture}(17,11)
\putps(0,-0.4)(0,-0.5){figRes}{figRes.ps}
\put(3.2,-0.5){\makebox(0,0){$1/\theta$}}
\put(8.9,-0.5){\makebox(0,0){$1/\theta$}}
\put(14.6,-0.5){\makebox(0,0){$1/\theta$}}
\end{picture}
\vspace{1cm}
\caption[SP]{\em Comparison of our method with lattice studies:
Diamonds denote the saddle-point action $S_k$
and squares the bubble-nucleation rate
$-\ln\left(I/m^3\right)$
for $k=1.2\,k_f$ and $2\,k_f$.
Dark circles denote the results for the nucleation rate from
the lattice study of ref.~\cite{twopone}.
Finally, the dashed straight lines
correspond to the action of
the saddle-point computed from the
potential of eq.~(\ref{eq:two20}).
\label{fig:Res}}
\end{center}\end{figure}
For $\lambda=0$
the potential of eq.~(\ref{eq:two20}) is unbounded from below and
the procedure we outlined in the previous paragraphs for matching
the theory simulated on the lattice may be problematic. However, one
may consider these results as applying to
the limit $\lambda\to 0$, so that no conceptual problems arise.
The values of $-\ln(I/m^3)$ computed at $1.2\,k_f$ and $2\,k_f$
are equal to a very good approximation, which
confirms the convergence of the expansion around the saddle-point and
the reliability of the calculation. The $k$ dependence of
the saddle-point action is cancelled by the prefactor, so that the
total nucleation rate is $k$ independent. Moreover, the
prefactor is always significantly smaller than the
saddle-point action.
The circles indicate the results of
the lattice simulations of ref.~\cite{twopone}.
The agreement with the lattice predictions is
good. More specifically, it is clear that the contribution of
the prefactor is crucial for the correct determination of the
total bubble-nucleation rate. Similar conclusions can be drawn
for $\lambda=0.1$ and $\lambda=0.2$.
For larger values of $\lambda$ the lattice simulations have been performed
only for $\theta$ significantly larger than 1. For smaller
$\theta$, nucleation events become too rare to be observable
on the lattice. As we discussed
earlier, the matching between the lattice and the renormalized actions
becomes imprecise for large $\theta$.
This indicates that we should expect
deviations of our results from the lattice ones, as the theory of
eq.~(\ref{renorm}) may be different than the simulated one.
These deviations start becoming apparent for the value $\lambda=0.25$, for
which the lattice simulations were performed with $\theta \sim 10$--20.
For $\lambda\geq 0.3$ the lattice results are in a region in which
the internal consistency criteria of our method for the reliability of
the expansion around the saddle-point are not
satisfied any more.
The consistency of our calculation is achieved
for $1/\theta \gta 0.12$ even for $\lambda=0.32$.
However,
the breakdown of the expansion around the saddle point is apparent
for $1/\theta \lta 0.12$.
The $k$ dependence of the predicted bubble-nucleation rate is
strong\footnote{The additional squares for $1/\theta=0.1$ correspond
to results from the numerical integration of eq.~(\ref{evpot}).}.
The prefactor becomes comparable to the saddle-point
action and the higher-order corrections are expected to be large.
The $k$ dependence of $S_k$ is very large. For this reason we
have not given values of $S_k$ in this case.
This behaviour
indicates that the field fluctuations become significant.
Typically, these fluctuations enhance the total rate and for
$1/\theta \lta 0.08$ can even compensate
the exponential suppression. Several similar examples
were given in refs.~\cite{first}--\cite{third}.
The reason for this behaviour can be traced to the form of the differential
operators in the prefactor (see eqs.
eqs.~(\ref{rrate})--(\ref{ak})).
This prefactor, before regularization, involves the ratio
$\det'(-\partial^2+U''_k(0)+W_{k}(r))/\det(-\partial^2+U''_k(0))$, with
$W_{k}(r)=U''_k\left(\phi_b(r)\right)
-U''_k(0)$. The function $W_{k}(r)$ always has a minimum away from
$r=0$ (see figs.~1c and 1g),
where it takes negative values. As a result the lowest
eigenvalues of the operator $\det'(-\partial^2+U''_k(0)+W_{k}(r))$ are smaller
than those of $\det(-\partial^2+U''_k(0))$. The elimination of
the very large eigenvalues from the determinants
through regularization does not affect this
fact and the prefactor $A_{k}$ is always larger than 1. Moreover,
in cases such as those depicted in fig.~2f it becomes exponentially large
because of the proliferation of low eigenvalues in
$\det'(-\partial^2+U''_k(0)+W_{k}(r))$.
In physical terms, this implies the
existence of a large class of field configurations of free energy comparable
to that of the saddle-point. Despite the fact that they are not
saddle points of the free energy
(they are rather deformations of such a point)
and are, therefore, unstable, they result in a dramatic increase of
the nucleation rate. This picture is similar to that of
``subcritical bubbles'' of ref.~\cite{gleiser}.
In ref.~\cite{nonpert}
the nucleation rate was
computed by first calculating a corrected potential that incorporates
the effect of such non-perturbative configurations. The pre-exponential factor
must be assumed to be of order 1 in this approach, as the effect of most
deformations of the critical bubble has already been taken into account in
the potential.
In our approach the non-perturbative effects are incorporated
through the prefactor.
Both methods lead to similar conclusions for the enhancement of
the total nucleation rate.
A final comment concerns the $\theta$ dependence of our results for
the nucleation rate in fig.~2. For a given value of $\lambda$, we observe a
linear dependence of $-\ln(I/m^3)$ on $1/\theta$
in the regions where the calculation is consistent. This is explained by the
fact that the dominant $\theta$ dependence of the potential
arises from the first term in the right-hand side of eq.~(\ref{incond}).
This is in agreement with the findings of ref.~\cite{twopone}.
In the latter
work, this linear dependence was not observed for $\lambda=0.3$ and 0.32.
The reason is that the range of $\theta$ used in the simulations was so large
that the first term in the right-hand side of eq.~(\ref{incond}) ceased to
be dominant. Our results for small $\theta$ display the expected behaviour
even for $\lambda=0.32$.
\setcounter{equation}{0}
\renewcommand{\theequation}{\thesection.\arabic{equation}}
\section{Conclusions}
In this paper we applied our approach for the calculation of
bubble-nucleation rates to (2+1)-dimen\-sional theories at non-zero
temperature. We studied these theories at coarse-graining scales $k$
below the temperature, where they display an effective two-dimensional
behaviour.
This provided the opportunity to check several points of our approach
that depend on the dimensionality of the system. For example,
the evolution equation for the coarse-grained potential
has a different form than the one in the effective three-dimensional
systems we studied in the past. Also, the
pre-exponential factor in the bubble-nucleation rate
has a logarithmic ultraviolet divergence before regularization,
instead of the leading linear divergence in three dimensions.
This divergence is reflected in the leading logarithmic dependence of
the regularized prefactor on the scale $k$
that acts as an ultraviolet cutoff (fig.~1).
More crucially, the complementarity between the $k$ dependence of
the prefactor and the saddle-point action was observed again.
This is a crucial point that guarantees that a physical quantity, such
as the bubble-nucleation rate, is independent of the scale $k$ that
we introduced as a calculational tool.
The validity of Langer's theory of homogeneous nucleation
was confirmed, as long as the prefactor gave a contribution to
the nucleation rate smaller than the leading exponential
suppression by the action of the saddle point.
Another important aspect of this work concerns the comparison
of our results with data from lattice simulations. This constitutes
a stringent quantitative test of our method. We found good agreement
between our results and the lattice data when
the renormalized action for the
theory that is simulated on the lattice is known (fig.~2). In
these cases, we first match the renormalized action and then
compute the nucleation rate.
For part of the range of the lattice parameters
this is not possible, because the renormalized action
cannot be obtained from the lattice action
perturbatively. However,
the internal consistency criteria of our method
provide a test of the reliability of our results in all cases.
\paragraph{Acknowledgements} We would like to thank
M. Gleiser and C. Wetterich for helpful discussions.
The work of N.T. was supported by the E.C. under TMR contract
No. ERBFMRX--CT96--0090.
\small
|
\section{Introduction}
This is the second of two papers describing the results of surveys
examining the properties of planetary nebulae (PNe) as observed in the
near-infrared ($\lambda = 1 - 2.5$ \micron). The first paper (Latter
et al.\ 1995; hereafter Paper I) presented an infrared imaging survey;
here we present the results of a near-infrared spectral survey.
There are several reasons why knowledge of the near-infrared (near-IR)
characteristics of PNe is important, as described in Paper I. In
order to interpret the imaging results and to learn more about the
physical conditions in the nebulae, the spectra of these objects must
be examined to understand the processes responsible for the
emission. There are many emission lines present in the 1 -- 2.5
\micron\ spectral region, most notably those due to recombination
lines of \ion{H}{1}, and lines of vibrationally excited H$_2$. Also
present are atomic lines of \ion{He}{1} and [\ion{Fe}{2}], and
emission from other molecular species such as CO and C$_2$. These
lines can act as diagnostic tools to probe the physical conditions
inside the nebula, sampling different regions and ranges of
temperature, density, and excitation than is seen by observing the
optical line emission. Some PNe also exhibit in the near-IR strong
continuum emission from hot dust. This emission becomes significant
longward of $\lambda = 2$ \micron\ in many of the PNe, requiring
near-IR spectroscopy to detect it and to differentiate between line
and continuum emission sources. Finally, the lower optical depth of
the PNe in the IR as compared to optical wavelengths allows us to
potentially see into regions of the nebula that are obscured by dust.
There have been several previous surveys that have explored the
properties of PNe in the infrared. Early spectroscopic and
photometric surveys (e.g., Gillett, Merrill, \& Stein 1972; Cohen \&
Barlow 1974) determined that there was an excess of IR emission over
what was expected from reflected continuum emission from the central
star. Other photometric surveys in the following years (Whitelock
1985; Persi et al.\ 1987) determined the primary sources of IR
emission to be stellar continuum, thermal dust emission, and thermal
\& line emission from the nebula itself. The near-IR color
characteristics of most PNe are unique and can be used to identify new
PNe and post-AGB objects (Garcia-Lario et al.\ 1990).
Recently there have been more detailed spectral observations of PNe in
the near-IR. Hrivnak, Kwok, \& Geballe (1994) surveyed a set of
proto-PNe in the H and K bands. In these objects the \ion{H}{1}
Brackett lines were observed in absorption, and most objects had CO
absorption or emission, indicating recent mass loss events. Rudy et
al.\ (1992, and references therein) and Kelly \& Latter (1995) have
surveyed several PNe and proto-PNe in the $\lambda = 0.5 - 1.3$
\micron\ range. Dinerstein \& Crawford (1998) have completed a survey
of a set of PNe in the K-band, focusing on excitation of molecular
hydrogen.
There are several unique aspects of the survey results presented here
that were made possible by the KSPEC spectrograph (see the instrument
description below). First, because of KSPEC's high sensitivity and
simultaneous sampling of the full near-infrared spectral range, we
were able to obtain data on a comparatively large number of objects
(41) in a short period of time. Using the relatively narrow and short
slit of the spectrograph, we sampled different regions of the PNe to
examine the emission throughout the nebula. In most of the other
surveys described above, larger beams were used that included much or
all of the object. Another aspect of the data presented here is
because of the cross-dispersed design of the instrument, the entire
$\lambda = 1.1 - 2.5$ \micron\ range is obtained at once, eliminating
the possibility of telescope pointing errors or other fluctuations
affecting the relative line strengths in the spectra. Finally, the
slit-viewing detector allowed precise positioning and guiding, so the
region of the PN being observed was well known for each spectrum. The
PNe observed in this survey were chosen to overlap with the near-IR
imaging survey (Paper I), along with several other optically-bright
PNe and unresolved objects that were not included in the imaging
survey.
\section{Observations and Data Reduction}
The observations were performed on several runs during the period 1992
October through 1994 September at the University of Hawaii 2.2m telescope
on Mauna Kea, using the near-IR KSPEC spectrograph (Hodapp et al.\ 1994).
KSPEC is a $\lambda = 1 - 2.5$ \micron\ cross-dispersed spectrograph that
has a separate slit-viewing IR array for acquiring the source and guiding.
The full spectral range is obtained in a single exposure, resulting in
accurate relative line measurements and highly efficient data acquisition.
The diffraction orders are well-matched to the atmospheric transmission
windows, with the K band in 3rd order (1.9 -- 2.5 \micron), H in 4th order
(1.45 -- 1.8 \micron) and J in 5th order (1.15 -- 1.32 \micron), with a
resolution R $\sim$ 700. The
1$\times$6 arcsecond slit provides a small aperture that was used to
sample different spatial regions of the nebula to search for spectral
variations.
Table 1 lists the nebulae observed and details of the observations.
Integration times for each exposure ranged from a few seconds for the
extremely bright sources to 5 minutes for faint sources. The off-axis
guider was used to keep a consistent on-source slit position. Multiple
``Fowler'' sampling was used to reduce the read noise. The number of
samples for a particular integration ranged from 4 to 16, with more
samples used for the longer integration times. The spectra were reduced
using IRAF; the extraction and processing of the spectral data were done
using the functions in the noao.twodspec and noao.onedspec packages.
Alternating source and sky integrations of the same length were taken and
differenced to remove sky and telescope background flux. Dome flats were
used to correct for pixel-to-pixel gain variations in the array. Stars of
known spectral type (either G0 or A0) were observed at the same airmass as
the nebulae immediately before and after the PNe observations and were
used for correction of the instrumental response and sky transmission.
Individual lines were removed from the stellar spectra, and then
normalized using a blackbody function of T = 5920 K for the G0 stars and T
= 10800 K for the A0 standards. The spectra were wavelength-calibrated
using observations of an Argon reference lamp. The wavelength values used
for lines greater than 1.1 \micron\ are from Rao et al.\ (1966); for lines
less than 1.1 \micron, the wavelengths were taken from Wiese, Smith, \&
Miles (1969) and corrected to vacuum wavelengths. The average 1$\sigma$
uncertainty in the measured wavelengths of the lines is about 5 \AA.
Infrared photometric standard stars were observed in the same way as the
PNe and used to flux calibrate the spectra. Absolute calibration is
difficult with these spectra because not all of the flux from the star
enters the narrow slit during a single integration, and the amount differs
for each integration depending on how well the star is centered on the
slit (the seeing at 2 \micron\ during typical observations was 0\farcs5 --
1\farcs0). The amount of light lost was estimated by the following
method: the full width at half maximum (FWHM) brightness of the standard
star was measured along the spatial direction of the slit, in the spectrum
with maximum flux for that star. It was then assumed that the point
spread function (PSF) is well represented by a two-dimensional Gaussian
distribution with the measured FWHM, and the amount of flux falling
outside of the slit was then calculated. This was typically 20 -- 30\% of
the total light for a single integration. The calibration for each star
was corrected by this factor, along with corrections for airmass.
Comparing results from different standard stars taken throughout the night
indicated that this method is accurate to approximately 20\%. For the
observations of the PNe, no correction was applied for the slit width or
length. The
length along the slit of the extracted regions for the three bands were
2\farcs0 (J), 3\farcs5 (H), and 4\farcs0 (K).
\section{Results }
The spectra are presented in Figures 1 -- 32. One to three PNe spectra
are plotted in each figure.
Tables 2 through 7 list the line identifications and extracted fluxes with
uncertainties for the spectra shown in the Figures.
The PN in this section are listed in Tables 2, 3, and 4.
There are a number of features not identified (indicated by question
marks in the tables). These features tend to appear above the
$\sim3\sigma$ level and must be considered real, though confirming spectra
would be valuable. Our search for possible identifications for these lines
has been careful, but perhaps not exhaustive. In addition to the
relatively low S/N of these lines, the moderate wavelength resolution is
not sufficient to differentiate between the several possible
identifications for each line.
The PNe spectra are separated into four groups that share common
characteristics. These are \ion{H}{1} recombination line-dominated,
\ion{H}{1} recombination line + H$_2$\ emission, H$_2$\ -- dominated, and
continuum-dominated. A fifth group of objects is included that contains
two objects that were at one time classified as PNe, but are now generally
regarded as being \ion{H}{2} regions (M 1--78 and K 4--45; Acker et al.\
1992). We do not discuss these objects further, but include them for
comparison. Within each group, the NGC objects are listed first, followed
by the remaining PNe in alphanumeric order. The morphological
classifications are given according to Balick (1987), unless otherwise
noted.
\subsection{\ion{H}{1} - line dominated}
The line emission in these PNe is dominated by lines of \ion{H}{1} and
\ion{He}{1}. In the J-band, the Paschen $\beta$ (Pa$\beta$) line is the most
intense, with contributions from lines of \ion{He}{1}, [\ion{Fe}{2}], and
\ion{O}{1}. In the H band, the Brackett series of \ion{H}{1} dominate,
with \ion{He}{1} emission at 1.7002 \micron\ and [\ion{Fe}{2}] emission at 1.6440
\micron\ present in some PNe. In the K band, the
brightest line is usually Brackett $\gamma$ (Br$\gamma$), with strong lines of
\ion{He}{1} at 2.058 and 2.112 \micron. When the \ion{H}{1} lines are
strong enough, one begins to see the Pfund series lines starting near 2.35
\micron\ where they are just beginning to be separated at this resolution.
There are also two unidentified lines at 2.199 and 2.287
\micron\ (Geballe et al.\ 1991) that appear in several PNe in this
category. A few of the spectra
shown here have contributions from central star continuum flux that is
larger towards shorter wavelengths, or warm dust continuum which is
stronger at longer wavelengths.
\subsubsection{NGC 1535}
NGC 1535 is classified as early round, and its near-IR spectrum is
dominated by emission lines of \ion{H}{1}. It has a bright ionized
shell of emission which is surrounded by a fainter halo (e.g., see
Schwarz, Corradi, \& Melnick 1992). Near-IR images were presented in
Paper I. This PN has previously been observed to have H$_2$\ lines in
absorption in the far-UV (Bowers et al.\ 1995). Recent observations
by Luhman et al.\ (1997) failed to detect H$_2$\ in emission in the $v = 1 \to 0$ S(1)
line. They attributed the earlier detection of absorption at shorter
wavelengths to the interstellar medium, or a region in the PN itself of
a much smaller size than the ionized zone.
The spectrum of NGC 1535 shown in Figure 1 was obtained at a position
centered on the brightest part of the ring directly west (W) of the
central star. We also fail to detect the H$_2$\ emission in the $v = 1 \to 0$ S(1)
line, at a 1 $\sigma$ level of about $5\times10^{-17}$ ergs cm$^{-2}$
s$^{-1}$ \AA$^{-1}$. There is some indication of emission from H$_2$\ in the $v
= 1\to 0$ Q(1) and $v=1\to 0$ Q(3) lines at the long wavelength end of
the spectrum. However, the spectrum is noisier in this region and
there is confusion with the Pfund-series \ion{H}{1} lines, therefore the H$_2$\
lines are not detected above the 3 $\sigma$ level.
\subsubsection{NGC 2022}
NGC 2022 is an early elliptical PN, but is morphologically very
similar to NGC 1535 in optical images (e.g., Schwarz et al.\
1992). The main difference between the two is a different relative
outer halo size as compared to the inner ring (the halo is relatively
smaller in NGC 2022). Zhang \& Kwok (1998) also find similar
parameters with their morphological fits of these two PNe. Near-IR
images of this PN were presented in Paper I.
NGC 2022 is also spectrally similar to NGC 1535, as seen in Figure
1. The spectrum of NGC 2022 was taken centered on the ring directly
east (E) of the central star. The dominant emission lines are those
of atomic hydrogen.
\subsubsection{NGC 2392}
The PN NGC 2392 (the ``Eskimo nebula'') is another double-shell nebula;
however, this PN has a significant amount of structure in the inner ring
and outer shell. Spectrophotometric (Barker 1991) and kinematic studies
(Reay, Atherton, \& Taylor 1983; O'Dell, Weiner, \& Chu 1990) that have
been carried out with optical imaging and spectroscopy have revealed the
abundances and ionization states and velocities of the various
components. Near-IR images of this PN were presented in Paper I.
The spectrum of NGC 2392 in Figure 1 was taken centered on the brightest
part of the ring directly E of the central star. This third early-type
round PN differs from the other two in Figure 1 primarily from the bright
\ion{He}{1} line at 2.058 \micron, and the [\ion{Fe}{2}] lines in the J
and H bands of the spectrum.
\subsubsection{NGC 3242}
NGC 3242 is an early elliptical with several interesting morphological
features. In addition to the bright elliptical ring, there are several
filaments and knots of emission in the central region, and two ansae that
are placed roughly along the major axis of the elliptical emission. Also,
there is a larger faint halo that envelopes the inner structure.
Spectra acquired at three different positions on the nebula are shown in
Figure 2, on the SE knot (NGC 3242SE), on the E section of the bright ring
(NGC 3242E), and on the SW halo (NGC 3242H). The spectra are similar in
all locations; the bright \ion{H}{1} lines are present in all positions,
along with stellar continuum at the shorter wavelengths. One difference
is that the \ion{He}{2} line at 2.189 \micron\ are much
brighter in the E ring than in the SE knot or halo position.
\subsubsection{NGC 6210}
NGC 6210 is a fairly compact PN with a core -- halo morphology similar to
other ellipticals. Phillips \& Cuesta (1996) performed a visible
wavelength spectroscopic study that revealed a complex velocity structure
which suggests multiple shells and possibly ``jets'' at various position
angles. The near-IR image presented in Paper I does not reveal much of
the structure. This PN is one of several in which Geballe et al.\ (1991)
detected unidentified emission at 2.286 \micron\ (but not at 2.199 \micron).
Three spectra are presented for this PN, and shown in Figure 3. They were
acquired with the slit centered on the core (Core), 1\arcsec\ east (E1),
and 3\arcsec\ east (E3). The core position shows a contribution from
stellar continuum in the $\lambda = 1 - 1.8$ \micron\ region that
decreases successively in the E1 and E3 positions. The unidentified feature at
2.287 \micron\ is detected at the E1 position, but not the 2.199 line,
similar to the findings of Geballe et al.\ (1991).
\subsubsection{NGC 6543}
NGC 6543 (the ``Cat's Eye'') has a complex morphology, with a high degree
of symmetry. Recent HST imaging (Harrington \& Borkowski 1994) has shown
more clearly the structure of rings, shock fronts, jets, and fast,
low-ionization emission-line regions (FLIERS) present in this PN.
The spectrum shown in Figure 4 was taken at the position S of the central
star and slightly E, where the two emission arcs cross and create a local
emission maximum (see Paper I, Figure 6a). The spectra is similar to the other PN in
the \ion{H}{1}/\ion{He}{1} - dominated class, and also show the
unidentified lines at 2.199 and 2.286 \micron.
\subsubsection{NGC 6572}
This young PN has a bipolar morphology in the near-IR, with its major axis
in the N-S direction, and a bright ring structure closer to the central
star (Paper I). The near-IR spectrum from 0.77 to 1.33 \micron\ was
measured by Rudy et al.\ (1991), and the UV and optical spectrum was
recently obtained by Hyung, Aller, \& Feibelman (1994), who found evidence
of variability.
Figure 5 shows two slit positions on the PN, one on the nebula center,
and one on the brightest location in the E lobe of the PN. Both the
core and E lobe spectra show a strong contribution from stellar
continuum flux. There is strong \ion{H}{1} and \ion{He}{1} line
emission, and relatively strong unidentified line emission at 2.199 and
2.286 \micron.
\subsubsection{NGC 6790}
This PN has been shown by radio continuum observations to be an
elliptically-shaped nebula with a diameter of roughly an arcsecond
(Aaquist \& Kwok 1990). Aller, Hyung, \& Feibelman (1996) obtained UV and
optical spectra and suggest that NGC 6790 is a relatively young object,
slightly more evolved than Hb 12. Kelly \& Latter (1995) obtained a 0.9
-- 1.3 \micron\ spectrum and find similarities to Hb 12 and AFGL 618.
The spectrum of NGC 6790 in Figure 4 shows a stellar contribution because
the slit was centered on the object, and includes the star as well as the
nebular emission that is typical of this class of PN.
\subsubsection{NGC 6803}
This compact elliptical PN has a uniformly bright disk with no
apparent structure in the optical, and is surrounded by a fainter halo
about twice the size of the bright shell (Schwarz et al.\ 1992). In the
KSPEC imaging channel, however, the PN was seen to be double-lobed,
with the lobes on the minor axis of the bright elliptical
region. Spectra were obtained at two positions, one centered on the E
lobe, and the other position 5 arcseconds NW in the halo region. The
lobe spectrum shows bright \ion{H}{1} and \ion{He}{1} lines, with a
stellar continuum contribution out to about 1.5 \micron. There is also
unidentified line emission at 2.286 \micron. The halo spectrum is
quite different, with weak Pa$\beta$\ and continuum emission, and is
probably reflected star and nebular emission.
\subsubsection{NGC 6826}
This elliptical PN is morphologically similar to NGC 3242, with a bright
elliptical inner ionized ring, ansae along the major axis, and a fainter
halo that envelopes the system (Balick 1987). The I-band image in Paper I
shows evidence for a shell between the inner bright ring and the outer
halo, and at the same radial distance as the ansae.
The spectra are shown in Figure 7. The bright core is dominated by
stellar continuum emission. The other positions show strong
\ion{H}{1} line emission. The spectrum labeled SW Lobe was taken on the inner
bright ring directly SW of the central star. The SW Halo position was
taken in the halo midway between the bright ring and the outer edge of the
PN. The lobe and halo emission is similar, except for relatively brighter
lines of \ion{He}{1} at 1.7002 \micron\ in the lobe. There is also some
continuum emission at the short wavelength end of the nebular positions,
which is probably scattered light from the central star.
\subsubsection{NGC 7009}
NGC 7009 (the ``Saturn'' nebula) is an elliptical with an interesting
twisted symmetry in its shell and in the various filaments and knots of
emission. Balick et al.\ (1998) recently published HST images that show
the ``microstructures'' in this PN. The images show that the inner knots
are actually groups of FLIERs, and jets in [\ion{N}{2}] are seen that
terminate at the tips of the nebula.
Two positions were sampled, one in the halo region on the W edge of the
PN, and one in the N part of the nebula. Both positions show a
stellar continuum emission contribution from the central star. The
halo emission is similar to the spectrum taken on the north edge of the
PN.
\subsubsection{NGC 7662}
NGC 7662 is a triple-shell elliptical, with a bright inner ring,
a fainter outer shell, and a very faint nearly circular halo (Hyung \&
Aller 1997). This PN also has a large number of complex microstructures,
recently examined using HST imaging by Balick et al. (1998). They suggest
that there is a prolate elliptical bubble around the central star
aligned perpendicular to the bright ring. The bright ring
is interpreted as a torus seen at roughly 30$^\circ$\ inclination.
The spectrum shown in Figure 9 was taken on the bright ring directly SE of
the central star. The N end of the slit was near the central star, so
some stellar continuum is seen in the spectrum, which is otherwise dominated
by \ion{H}{1} and \ion{He}{1} lines.
\subsubsection{IC 351}
This compact PN has a double-lobed structure with a round halo (Hua \&
Grundseth 1986; Aaquist \& Kwok 1990; Manchado et al.\
1996). Feibelman, Hyung, \& Aller (1996) obtained UV and visible light
spectra of IC 351 that show it to be a high excitation nebula, but
without the presence of the usual silicon lines, and suggest that the
silicon atoms could be locked up in grains.
Our spectrum (Figure 9) taken centered on the PN suffers somewhat from an incomplete
subtraction of OH airglow lines, due to the sky frames not being taken
properly for the on-source images. The OH lines show up in emission
mainly in the H and K-band portions of the spectrum. However, the
major features of \ion{H}{1} and \ion{He}{1} emission lines can be
seen.
\subsubsection{IC 418}
The spectrum of this well-studied young, low-excitation PN was
previously shown to be dominated by lines of \ion{H}{1} and
\ion{He}{1} in the near-IR, with a hot dust continuum (Willner et
al.\ 1979; Zhang \& Kwok 1992; Hodapp et al.\ 1994). Hora et
al.\ (1993) and Paper I showed broad- and narrow-band near-IR images of
the PN, showing the elliptical, double-lobed structure in the IR.
Three positions in the nebula were observed to determine the spectral
variations across the object. The positions observed were on the central
star, the peak of the E lobe, and in the E halo region outside of the bright
ring (Figure 10). In the central position, stellar continuum is visible,
rising towards shorter wavelengths. The nebular lines are similar in the
central and lobe positions. The halo emission is almost devoid of lines;
there is some faint Pa$\beta$\ present as well as Br$\gamma$. This is possibly
reflected from the bright lobes. The main component of the halo emission
is a weak continuum that rises toward longer wavelengths.
\subsubsection{IC 2149}
This peculiar PN has a bright core and a roughly bipolar nebula
extending approximately E-W, but does not show the usual H$_2$\ signature
of bipolar PN. The two spectra shown in Figure 11 were taken centered
on the bright central star, and on the E lobe. The core spectrum shows strong
stellar continuum, with the nebular lines superimposed. The E lobe
emission is primarily from lines of \ion{H}{1} and \ion{He}{1}, in
addition to weak continuum emission which is probably reflected from
the central star.
\subsubsection{IC 3568}
This round PN consists of spherical shells, an inner bright one and a
outer halo. Balick et al.\ (1987) showed that the structure was
consistent with simple hydrodynamic models of PN that are shaped by
interior stellar winds. The spectrum taken on the N edge of the PN
is shown in Figure 12. The predominant features are \ion{H}{1}
and \ion{He}{1} emission lines, and low level continuum emission which is
probably reflected from the central star.
\subsubsection{IC 4593}
Bohigas \& Olguin (1996) obtained spectroscopy and imaging of this PN
which has two inner shells surrounded by an outer highly excited halo. IC
4593 is unusual in that the condensations outside of the inner region are
located asymmetrically in the SW region.
Two spectra were taken on this PN, one positioned on the central star, and
the other at a position 3\arcsec\ E (Figure 12). The core shows bright
stellar continuum, with nebular lines superimposed, and Pa$\beta$\ absorption.
The E spectrum has the typical \ion{H}{1} and \ion{He}{1} emission lines.
The sky subtraction was not of high quality for this spectrum, which
resulted in OH airglow lines showing up in emission in the H and K
spectral regions.
\subsubsection{J 320}
Images of J 320 (Balick 1987) show it to have a central region that is
elongated roughly E-W, but the low level flux has a N-S elongation
indicating a shell or streamer extending in this direction. Three spectra
for J 320 are shown in Figure 13, centered on the core (C), offset
1\arcsec\ N, and offset 1\farcs5 N. The core has a stronger
contribution from stellar continuum, but otherwise the spectra are
similar. We therefore detect no spectral differences between the N extension
and the nebula near the core.
\subsubsection{M 4--18}
This young, low-excitation PN was recently imaged with HST by Dayal et
al.\ (1997) and shows a toroidal shell surrounding the central star. There
is strong mid-IR emission from warm ($\sim$ 200 K) dust that has a similar
morphology but the position angle of the dust emission maxima is
orthogonal to those shown in H$\alpha$ emission. The spectrum of the core
of this PN shown in Figure 14 has strong stellar continuum, as well as
\ion{H}{1} and \ion{He}{1} line emission. The slit was positioned N-S
across the compact ring of the PN, so the ionized regions of the nebula
are included in this spectrum.
\subsection{\ion{H}{1} - line and H$_2$\ emission}
The emission lines present in the spectra of this PNe group contain those
mentioned in the previous section plus lines of molecular hydrogen. The
H$_2$\ lines are strongest in the K band, although in some objects there are
lines visible in the H and J bands as well. In the objects where more
than one slit position was measured, there is often a large change in the
relative line strength of \ion{H}{1} versus H$_2$\ emission, indicating that
the emission is being produced in different regions of the nebula. In
general, the H$_2$\ emission is more likely to be in the outer regions of the
PNe, whereas the \ion{H}{1} emission lines more closely trace the ionized
regions and has similar morphology to the visible appearance. The PNe in
this group all have some bipolar symmetry in their shape, most being of
the ``butterfly'' morphology characterized by narrow equatorial regions and
large bipolar lobes (e.g., M 2--9, Hb 12). However, some are classified as
elliptical based on the shape of the brightest components. Tables 5 and 6
list the line identifications and extracted fluxes for the PNe in this
section.
\subsubsection{Molecular hydrogen excitation in PNe}
A near-infrared H$_2$\ emission line spectrum can occur through slow
electric quadrupole vibration-rotation transitions in the ground
electronic state. Because the allowed transitions are such that $T_{\rm
ex} \gtrsim 1000$ K is required to produce a detectable near-IR H$_2$\ spectrum,
special excitation conditions must exist when the near-IR
spectrum is present. In Paper I we discussed mechanisms of H$_2$\
excitation in PNe; see also Kastner et al.\ (1996). If detected in
sufficient number, the observed H$_2$\ line ratios are an excellent
diagnostic for determining the relative importance of shocks and UV
photons in a photodissociation region (PDR) for the excitation of the H$_2$\
emission. Even if the excitation mechanism cannot be determined, the
presence of H$_2$\ emission is important to understanding the conditions in
PNe and how they evolve through wind interactions and photodissociation.
An analysis of near-IR H$_2$\ emission can determine an ortho-to-para
(O/P) ratio, the rotational excitation temperature $T_{ex}(J)$, and
the vibrational excitation temperature $T_{ex}(v)$ of the
molecules. If the rotational and vibrational excitation temperatures
differ, then UV excitation is indicated. This is most readily
determined by comparing the column densities in the upper state
vibration-rotation levels with the upper state energy (in temperature
units; see Hora \& Latter 1994, 1996 for a full discussion; see also
Black \& van Dishoeck 1987; Sternberg \& Dalgarno 1989). When many
H$_2$\ lines are detected, especially those from highly excited levels
that fall in the J band, a much stronger case can be made for the
importance of UV excitation than does the traditional $v = 2\to 1$
S(1) to $v = 1 \to 0$ S(1) line ratio (e.g.\ Hora \& Latter 1994, 1996). The
comparison of the column densities to the upper state energy levels
has been done for the PNe with detected H$_2$\ emission and the results
are summarized in Table 8. Only the rotational excitation
temperatures are listed. For collisionally- (shock-) excited spectra,
the rotational and vibrational excitation temperatures are coupled and
the same. For UV excited spectra, the vibrational temperature is the
result of a cascade through levels and not a thermal process. The
rotational levels are easily thermalized by collisions. The observed
O/P ratio is in general rather uncertain, especially if only $v = 1$
lines are detected. We did not attempt to determine the O/P ratio for
objects without sufficient line detections. An observed O/P ratio
lower than 3 indicates that the H$_2$\ emission is not thermally
excited. A subthermal O/P ratio as determined from near-IR spectra is
not caused by the UV excitation process itself, but is a function of
chemistry and density in the PDR (e.g.,\ Hora \& Latter 1996; Black \&
van Dishoeck 1987), and might not be indicative of the ``true'' O/P
abundance ratio of the H$_2$\ (see Sternberg \& Neufeld 1999). Selected
excitation diagrams for several objects that are discussed below are
shown in Figure 33.
\subsubsection{NGC 40}
The morphological classification of NGC 40 is middle elliptical (Balick
1987), which seems to contradict the previously observed strong correlation
between bipolar morphology and H$_2$\ detection. However, if one examines
the low-level emission in the N and S regions of this PN (see Paper I),
one can see material that has broken through and expanded beyond the
elliptical shell defined by the E and W bright lobes. Mellema (1995) has
found the morphology consistent with models of ``barrel''-shaped PNe, which
have roughly cylindrical emission regions slightly bowed outwards at the
equatorial plane, and less dense polar regions. Higher resolution and
more sensitive optical imaging has recently been carried out by Meaburn et
al.\ (1996) show gas escaping from the polar regions of the PN, with other
filamentary structure in the outer halo.
This is the first reported detection of H$_2$\ in NGC 40. The spectrum was
taken centered on the W lobe, and the H$_2$\ lines are relatively weak
compared to the \ion{H}{1} and \ion{He}{1} lines from the ionized gas in
this region. The H$_2$\ emission was not detected in the narrowband imaging
surveys of Paper I or Kastner et al.\ (1996), so the molecular emission
must be confined to a region near the bright ionized gas that dominates
the spectrum.
The data suggest that the H$_2$\ is
shock-excited. However, insufficient line detections make this result less
than firm.
\subsubsection{NGC 2440}
NGC 2440 is a bipolar PN with complex morphological and spectral
structure. In the optical, the nebula is bipolar with the major axis in
roughly the E-W direction for the large outer lobes (Balick 1987; Schwartz
1992). However, there are two bright lobes near the core that are
positioned along an axis roughly perpendicular to the major axis of the
outer lobes. There are two fainter knots that are also along a roughly
E-W axis, but not aligned with the outer lobes. There are filaments and
knots throughout the lobes. Lopez et al.\ (1998) finds up to three
outflowing bipolar structures in the lobes, and find from their kinematic
study that the inner bright lobes (their lobes ``A'' and ``B'') are the
emission maxima from a radially-expanding toroid viewed nearly in the
plane of the sky.
In the near-IR, the inner pairs of lobes are also prominent, but the large
E-W lobes are not visible (see Paper I). Instead, there is a circular
outer halo visible in H$_2$\ that is not quite centered on the inner lobe
structure. Also visible are faint H$_2$\ ``spikes'' that extend from the
center to the circular outer halo, roughly in the equatorial plane of the
large optical E-W lobes (Latter \& Hora 1998).
The spectra shown in Figures 15 and 16 were taken at three
different positions in
the PN: on the N lobe (of the innermost bright pair of lobes),
on the fainter E knot, and on a clump of H$_2$\ emission located on the
NE edge of the outer circular halo (see Paper I, Figure 4a; it is the clump
visible at the upper left corner of the ``H$_2$\ sub'' image). The N lobe
exhibits \ion{H}{1} and \ion{He}{1} lines from the ionized gas in this
region, but also has significant H$_2$\ emission. There is also strong
[\ion{Fe}{2}] emission at 1.64 and 1.257 \micron. The E knot also displays
similar \ion{H}{1}, \ion{He}{1}, [\ion{Fe}{2}], and H$_2$\ emission,
although fainter.
In contrast to the inner regions, the NE clump spectrum in Figure 16
is dominated by H$_2$\ emission, with the only \ion{H}{1} lines detected
being Pa$\beta$\ and Br$\gamma$. There is strong [\ion{Fe}{2}] emission at 1.64 and
1.257 \micron\ in this region as well. The excitation analysis for
the three positions observed showed that they are
UV-excited, except for the E knot
position for which there is insufficient data. The low value of the
observed O/P ratio is suggestive of the H$_2$\ emission arising from a PDR
at this location as well.
Since the inner region of NGC 2440 is morphologically complex and any line
of sight through the PN is likely to intersect several distinct regions,
it is probably the case that the ionized and molecular zones are not mixed
as the spectra might seem to indicate, but that the slit simply includes
several nebular components, or is looking through a PDR and is sampling
both the molecular and the recently ionized gas.
\subsubsection{NGC 6720}
NGC 6720 (the ``Ring Nebula'') is probably the best-known PN, and is the
archetype for the ring or elliptical morphology that characterizes the
brightest part of the nebula. The emission is not consistent with a
uniform prolate shell, however, since the ratio of flux between the edge
and center of the ring is higher than expected from a limb-brightened
shell (Lame \& Pogge 1994). Balick et al.\ (1992) have suggested that NGC
6720 is actually a bipolar PN viewed along the polar axis, based on
narrow-band imaging and high-resolution spectroscopic observations. This
view is supported by the presence of H$_2$\ in the nebula and halo, which
correlates strongly with bipolar morphology. Guerrero, Manchado, \& Chu
(1997) draw different conclusions, however, based on their chemical
abundance and kinematic study of the nebula. They argue that the Ring has
a prolate ellipsoid structure, with a halo of remnant red giant wind.
Our spectra of the Ring (Figure 17) were obtained at two positions, one on
the bright ring directly N of the central star, and the second position
several arcseconds further north, off the bright ring but on a moderately
bright (in H$_2$) position in the halo. Both positions show bright H$_2$\
emission, with the lobe position also showing contributions from emission
lines of \ion{H}{1} and \ion{He}{1} from the ionized gas, as one would
expect based on the visible wavelength and IR images showing the
distribution of the line emission. The ring spectrum is strongly UV
excited, indicating it is the PDR interface to the outer molecular shell.
\subsubsection{NGC 7026}
The late elliptical PN NGC 7026 has two bright lobes on either side (E-W)
of the central star, with fainter bipolar emission extending roughly N-S
from the core. Cuesta, Phillips, \& Mampaso (1996) obtained optical
spectra and imaging of this object and found kinematically complex
structure, with several separate outflows at the outer edge of an inner
spherical shell, and suggested that the primary shell may be undergoing
breakup in transition to a more typical bipolar outflow structure.
Two positions were sampled in NGC 7026, shown in Figure 18, centered on
the E and W bright lobes near the central star. The lobe spectra are
nearly identical, as one might expect from the symmetry in this PN. The
H$_2$\ emission in this PN is fairly weak at these positions. This might be
due to the H$_2$\ being concentrated in other regions of the PN, and not in
the bright ionized lobes that were sampled by the spectra presented here.
We are unable to determine the excitation mechanism.
\subsubsection{NGC 7027}
NGC 7027 is one of the most highly studied PN at all wavelengths,
particularly in the infrared because of its brightness and wealth of
spectral features. Treffers et al.\ (1976) obtained a spectrum for
$\lambda = 0.9 - 2.7$ \micron\ with a beam that included the entire
nebula. They identified the major near-IR spectral components, including
the first detection of H$_2$\ lines in a PN, and the first detection of the
unidentified line at 2.29 \micron. Since then, several near-IR spectra
have been published, including Scrimger et al.\ (1978), Smith, Larson, \&
Fink (1981), Rudy et al.\ (1992), and Kelly \& Latter (1995).
The spectra shown in Figure 19, one taken centered on the W bright lobe,
and the other at the brightest position in H$_2$\ of the NW lobe (see
Paper I). Both show \ion{H}{1} and H$_2$\ emission; the H$_2$\ is relatively
stronger in the NW position than in the bright lobe. Narrowband imaging
has shown that the \ion{H}{1} and \ion{He}{1} emission is primarily in the
bright inner ring of the nebula, and the H$_2$\ emission is in what appears
to be bipolar lobes outside of this shell (Graham et al.\ 1993a,b; Paper
I; Latter et al.\ 1998). It has been argued before based on morphology
that the H$_2$\ is in a PDR (Graham et al.\ 1993a). Our data clearly
demonstrate this to be the case, with the H$_2$\ showing a strongly UV
excited spectrum in a relatively high density medium (see Figure 33b).
\subsubsection{BD+30$^{\circ}$3639}
The young PN BD+30$^{\circ}$3639 is well-studied in the infrared,
and is remarkable primarily because of its
large IR emission excess. It has many similarities to NGC 7027, including
its IR morphology and the presence of H$_2$\ in the near-IR and
unidentified IR (UIR) emission features in the mid-IR spectrum,
which are usually attributed to
polycyclic aromatic hydrocarbons (PAHs).
Rudy et al.\ (1991) obtained a $\lambda = 0.46 -
1.3$ \micron\ spectrum of BD+30$^{\circ}$3639; high-resolution visible and near-IR
images were recently obtained by Harrington et al.\ (1997) and Latter et
al.\ (1998), and ground-based near- and mid-IR images have been presented
by Hora et al.\ (1993), Paper I, and Shupe et al.\ (1998).
Three positions in BD+30$^{\circ}$3639 were sampled in the spectra presented in Figure
20; the emission peak on the N lobe of the ring, the E side of the ring,
and on the H$_2$\ emission region located approximately 3\arcsec\ E of the
ring. These spectra show a steady progression of decreasing emission from
the ionized gas and increasing molecular emission as one moves east. As
for NGC 7027, the H$_2$\ emission in BD+30$^{\circ}$3639
is UV excited (Figure 33a) and defines the
PDR (see also Shupe et al.\ 1998).
\subsubsection{Hubble 12}
Hubble 12 (Hb 12) has been notable primarily because it represents one of
the clearest cases known of UV excited near-IR fluorescent H$_2$\ emission
(Dinerstein et al.\ 1988; Ramsay et al.\ 1993). Our Hb 12 results from
this survey and our imaging survey were presented in a previous paper
(Hora \& Latter 1996); the spectra are reproduced here for comparison with
the rest of the survey.
Dinerstein et al.\ had mapped the inner structure and found it to be
elliptical surrounding the central star; our deep H$_2$\ images showed the
emission to be tracing the edges of a cylindrical shell around the star,
with faint bipolar lobes extending N-S. We also detected [\ion{Fe}{2}]
line emission at 1.64 \micron\ in a position along the edge of the shell.
The H$_2$\ line ratios observed were in excellent
agreement with predictions by theoretical H$_2$\ fluorescence calculations,
and no significant differences were found between the excitation in the
two positions of the nebula that were sampled (see also Luhman \& Rieke
1996).
\subsubsection{IC 2003}
IC 2003 is a round, high-excitation PN that has a ring of emission, with a
bright knot on the S edge (Manchado et al.\ 1996; Zhang \& Kwok 1998).
Feibelman (1997) obtained IUE spectra of this PN that shows a wealth of
nebular and stellar lines. The IR spectra presented in Figure 22 taken
in the center of the PN shows
that there is little continuum from the nebula; the emission is primarily
from lines of \ion{H}{1} in the J, H, and K bands. There is strong unidentified
emission at 2.286 \micron\ but none detected at 2.199 \micron. H$_2$\ emission
is tentatively detected in the K-band, in the $v = 1 \to 0$ S(1), $v=3\to 2$ S(1), and
$v=1\to 0$ Q(1) lines. Each of the lines are detected at roughly a
2$\sigma$ level. The line fluxes are not reliable or numerous enough to
allow fitting of the line ratios.
\subsubsection{IRAS 21282+5050}
The young, carbon-rich PN IRAS 21282+5050 has been
identified as having an 07(f)-[WC11] nucleus (Cohen \& Jones 1987) with
possibly a binary at its center. Strong $^{12}$CO has been detected in a
clumpy expanding shell (Likkel et al.\ 1988) with elongated emission N-S.
Shibata et al.\ (1989) believe the elongated emission suggests the
presence of a dust torus in the E-W direction; however, Meixner et al.\
(1993) was evidence for a clumpy, expanding elliptical envelope. The
elongated structure is also seen in the visible (Kwok et al.\ 1993). Weak
continuum flux at 2 and 6 cm suggests a young PN just beginning to be
ionized (Likkel et al.\ 1994; Meixner et al.\ 1993). Kwok et al.\ (1993)
believe there has been a recent sharp drop in luminosity based on the
measured CO/FIR ratio. Weak HCO$^+$ and $^{13}$CO are present (Likkel et
al.\ 1988).
Two positions were sampled on IRAS 21282+5050, centered on the bright core, and
offset approximately 3\arcsec\ N and 3\arcsec\ W. The spectrum of the offset
position is shown in Figure 22. The core is dominated
by continuum emission from the central star. Also present are both
emission lines from the ionized gas, and H$_2$\ features in the K-band.
The nebula is compact, about 4\arcsec\ in diameter at K (Paper I). The slit
therefore samples a slice through the entire nebula, and as a result this
spectrum does not necessarily imply that the molecular and ionized gas is
mixed. The Lobe spectrum shows primarily lines of H$_2$\ (with the OH night
sky lines showing up in absorption because of imperfect sky subtraction in
this spectrum). The lack of emission lines due to \ion{H}{1} and
\ion{He}{1} in the Lobe spectrum indicates that the H$_2$\ emission is
predominantly in the outer regions of the PN. The data are suggestive of
shock excitation, but this should be considered tentative.
\subsubsection{M 1--16}
M 1--16 is a PN with a near-IR bright central region and bipolar lobes with
fast winds extending at least 35\arcsec\ from the core (Schwartz et al.\
1992; Aspin et al.\ 1993; Sahai et al.\ 1994). Several spectra were
obtained in this PN scanning across the central region; the two positions
shown in Figure 23 are on the core position and 1\arcsec\ S of the core.
Both positions show H$_2$\ emission; the S position is slightly brighter in
both H$_2$\ and the ionized nebular lines. Our data reveal that the H$_2$\ is
UV excited in both regions observed. This had be suggested earlier by
Aspin et al.\ (1993). The core shows a slight rise towards long wavelengths
indicating emission from warm dust continuum.
\subsubsection{M 1--92}
M 1--92 (``Minkowski's Footprint'') is a bipolar proto-planetary nebula
similar in near-IR appearance to AFGL 618, and has evidence of highly
collimated outflows along the bipolar axis (Paper I; Trammell \& Goodrich
1996 and references therein). Two positions were sampled in M 1--92, one in
the core and one on the NW bipolar lobe. There are problems with the sky
background subtraction in both spectra, which are most prominent in the
$\lambda = 1.9 - 2.1$ \micron\ region of the spectrum, but also contribute
to a lower signal to noise ratio (S/N) over the whole dataset.
Nevertheless, the primary characteristics are apparent. The core region
is dominated by strong warm dust continuum emission. There is also weak
Br$\gamma$\ and Pa$\beta$\ emission, but the other most other \ion{H}{1} and
\ion{He}{1} lines are too weak to be detected. The lobe position shows
weak H$_2$\ emission. The emission appears to be shock-excited, but the low
excitation suggested by our data is suggestive of UV excitation. Data of
higher S/N are required to discern the dominant excitation mechanism.
There might also be H$_2$\ emission near the core that is being masked by
the strong continuum emission. In both positions, there also seems to be
CO bandhead emission at $\lambda = 2.3 - 2.5$ \micron\, although the S/N
is not high in these regions.
\subsubsection{M 2--9}
M 2--9 (the ``Butterfly'') is a highly symmetric bipolar nebula, with lobes
extending from opposite sides of a bright central core, nearly in the
plane of the sky. Bright knots of emission are visible in the lobes at
the N and S ends. Our results for M 2--9 from this survey were previously
presented in Hora \& Latter (1994), and some of the spectra are reproduced
in Figures 25 and 26 for comparison. High-resolution imaging in several
near-IR lines indicated that the lobes had a double-shell structure, with
the inner shell dominated by \ion{H}{1} and \ion{He}{1} line emission from
ionized gas and continuum emission scattered from the central source, and
the outer shell (the ``Lobe O'' spectrum in Figure 25)
of the lobes showing strong H$_2$\ emission which exhibit a
spectrum consistent with UV excitation in a PDR. The core region shows a
strong dust continuum component, as well as emission lines of
\ion{H}{1}, \ion{He}{1}, \ion{Fe}{2}, [\ion{Fe}{2}] and \ion{O}{1}. The N
knot has strong [\ion{Fe}{2}] emission, with relatively weaker \ion{H}{1},
\ion{He}{1}, and H$_2$\ emission.
\subsubsection{Vy 2--2}
Vy 2--2 is a compact PN, so very little is known about its morphology. The
spectrum obtained in this survey was taken centered on the bright core and
the slit sampled most or all of the emission from this object. The
spectrum contains stellar continuum, lines of \ion{H}{1} and \ion{He}{1}
emission from the nebula, and weak H$_2$\ emission. This detection confirms
the indication of H$_2$\ emission as reported by Dinerstein et al.\ (1986).
The spectrum shown in Figure 26 is similar to others in this category,
such as BD+30$^{\circ}$3639 and NGC 2440, where several nebular components are
superimposed because of the position and size of the slit. As for those
objects, the H$_2$\ spectrum in Vy 2--2 is also UV excited.
\subsection{H$_2$\ dominated}
The PNe in this group have spectra that primarily contain emission lines
of H$_2$. These objects all have bipolar morphology, and most are young or
proto-PNe (PPNe). The PPNe also have warm continuum dust emission or
stellar continuum that is strongest in the core. Table 7 lists the line
identifications and fluxes for the PNe in this section.
\subsubsection{NGC 2346}
NGC 2346 is a PN with faint bipolar lobes seen clearly in H$_2$\ emission
(e.g.,\ Paper I). The brightest part of the nebula is in the ``equatorial''
region near the central star where the bipolar lobes meet. Walsh,
Meaburn, \& Whitehead (1991) performed deep imaging and spectroscopy that
showed the full extent of the lobes, and they model the PN as two
ellipsoidal shells that are joined near the central star. The distribution
of the H$_2$\ emission is similar to the optical (Zuckerman \& Gatley 1988;
Kastner et al.\ 1994; Paper I).
The near-IR spectrum of NGC 2346 in Figure 27 is dominated by UV-excited
H$_2$\ emission, as shown in Figure 33d. The spectrum was obtained with the
slit positioned on the bright condensation to the W of the central star.
There is also weak Pa$\beta$\ and Br$\gamma$\ emission seen, which is possibly
reflected from near the central star.
\subsubsection{J 900}
The PN J 900 is a bipolar nebula with an unusual ``jet''-like structure
and an outer shell structure that is seen primarily in H$_2$\ emission
(Shupe et al.\ 1995; Paper I). The spectrum of J 900 shown in Figure 27
was obtained at a position N of the brighter lobe just NW of the central
star, centered on the ``jet'' of emission. Problems with sky-subtraction
caused the J and H-band portions of the spectrum continuum to be slightly
negative. There is no detected continuum in any part of the spectrum. The
H$_2$\ spectrum is shock-excited in a moderate velocity wind (Figure 33e).
\subsubsection{AFGL 618}
AFGL 618 is a carbon-rich, bipolar reflection nebula with a relatively hot
central star ($\approx 30,000$ K), similar spectra in the two lobes, and
the eastern lobe is significantly brighter than the other. In this as in
other ways, the object bears a great resemblance to AFGL 2688 (see below),
despite the fact that their central star temperatures differ by about a
factor of 5. The visible spectrum shows numerous emission lines
characteristic of
ionized gas (Westbrook et al.\ 1975; Schmidt \& Cohen 1981)
which are scattered by dust into the line of sight, with a small
\ion{H}{2} region surrounding the central object (Carsenty \& Solf 1982;
Kelly, Latter, \& Rieke 1992). The near-IR spectrum of AFGL 618 is also
dominated by rotation-vibration lines of H$_2$\ (Thronson 1981, 1983; Latter
et al.\ 1992; Paper I).
AFGL 618 exhibits a rich spectrum of molecular line emission (Lo and
Bechis 1976; Knapp et al.\ 1982; Cernicharo et al.\ 1989; Kahane et al.\
1992; Martin-Pintado \& Bachiller 1992; Bachiller et al.\ 1997; Young
1997). The lines detected include $^{12}$CO, $^{13}$CO, C$^{17}$O,
C$^{18}$O, CS, NH$_3$, HCN, HCO$^+$, CN, and \ion{C}{1}.
Two of the positions sampled are presented here in Figure 28 -- the core
spectrum and one taken 2\farcs4 E of the core. Both spectra show strong
H$_2$\ emission, along with [\ion{Fe}{2}] and weak Pa$\beta$\ and Br$\gamma$. In
addition, the core has a warm dust continuum that is apparent throughout
the spectrum, and clear CO bandhead features in the 2.3 -- 2.4 \micron\
region. The CO features are also present but at lower levels in the
2\farcs4 E spectrum position. Our analysis of the H$_2$\ spectrum confirms
the earlier results by Latter et al.\ (1992) -- the spectrum is dominated by
a shock-heated component, but a UV excited component is clearly present as
well (Figure 33c).
\subsubsection{AFGL 2688}
AFGL 2688 (the ``Egg Nebula'') is a bipolar reflection nebula (Ney et al.\
1975) at visible and near-infrared wavelengths. It has a central star that
exhibits the spectrum of a carbon-rich supergiant (Crampton et al.\ 1975; Lo
\& Bechis 1976). Similar in visible appearance to AFGL 915, each lobe
shows two ``jets'' or ``horns'' extending away from the central region
(Crampton et al.\ 1975; Latter et al.\ 1993; Sahai et al.\ 1998a). The lobes
have identical spectra at visible wavelengths, but their brightness
differs significantly (Cohen \& Kuhi 1977). The near-IR spectrum is
dominated by H$_2$\ rotation-vibration lines (Thronson 1982; Beckwith 1984;
Latter et al.\ 1993). A central source is seen in the mid-IR and longer
wavelengths, with fainter extended emission along the axis of the nebula
(Hora et al.\ 1996). There is an enigmatic equatorial region seen in H$_2$\
emission and might be traced by other molecular species, such as HCN
(Latter et al.\ 1993; Bieging \& Ngyuen-Quang-Rieu 1996; Sahai et al.\ 1998b).
Similar to AFGL 618, this object also has a rich molecular content.
SiC$_2$ is seen in absorption (Cohen \& Kuhi 1977); this feature is
usually found in stars of the highest carbon abundance. Strong absorption
features of C$_3$ and emission in C$_2$ (Crampton et al.\ 1975) are
present, while C$_2$ is also seen in absorption in reflected light from
the lobes (Bakker et al.\ 1997). The CO $J = 1 \to 0$ line shows three
distinct velocity structures (Kawabe et al.\ 1987; Young et al.\ 1992).
Our results for this object from this survey were previously presented in
Hora \& Latter (1994, 1995) and our narrowband imaging in Latter et al.\
(1993). The spectra are reproduced here for comparison with the rest of
the survey. Spectra were obtained at several positions in the nebula,
including positions along the N lobe, and in the equatorial region (see
Hora \& Latter 1994 for details). The emission is segregated; the core is
dominated by continuum emission, there are emission lines of C$_2$ and
CN further from the core along the lobes, and the H$_2$\ emission is
confined to the ends of the lobes and in the equatorial region in what
appears to be a ring or toroidal structure (Latter et al.\ 1993; Sahai et al.\
1998b). Our analysis of the H$_2$\ line ratios showed that the emission is
collisionally excited in shocks, with no discernible difference between
the emission in the lobes and the equatorial region.
\subsection{Continuum - dominated}
These are young PNe or PPNe that have strong warm dust continuum and
little line emission. The strongest component is in general the core,
with most of the emission from an unresolved point source. In some of the
nebulae, emission structure extends a few arcseconds from the core region.
Also, in objects such as AFGL 915, they are associated with larger optical
nebulae that extend arcminutes from the core. In this survey, only the
regions near the core were sampled.
\subsubsection{AFGL 915}
AFGL 915 (the ``Red Rectangle'') is a carbon-rich biconical reflection
nebula with a metal-depleted spectroscopic binary at its center (Cohen et
al.\ 1975). The nebula appears axially symmetric and shows spikes running
tangent to the edge of the bicone. Surrounding the post-AGB star at its
center is a circumbinary disk viewed edge-on (Jura, Balm, \& Kahane 1995)
which could be oxygen-rich (Waters et al.\ 1998).
C$_2$ and CN are not detected near the binary, though C$_2$ is present in
emission in the reflection lobes. CH$^+$ (0,0) and (1,0) are detected in
emission (Bakker et al.\ 1997; Balm \& Jura 1992). CO is underabundant,
with relatively weak emission and broad wings detected (Dayal \& Bieging
1996; Greaves \& Holland 1997; Loup et al.\ 1993; Bujarrabal et al.\
1992). Glinski et al.\ (1997) found CO and \ion{C}{1} in the UV in both
absorption and emission. They expect strong CO overtone emission in the
IR based on their observations of hot CO emission and absorption in the
UV. The object shows ERE (extended red emission) from about $\lambda =
5400$ to 7200 \AA\ and a set of emission bands around 5800 \AA\ (Schmidt,
Cohen, \& Margon 1980) whose carriers might be the same material as the
carriers of the DIBs (diffuse interstellar bands). This object also shows
strong emission in the PAH bands at 3.3, 7.7, and 11.2 \micron\ (Cohen et
al.\ 1975), which are located predominantly in the lobes and spikes of
emission (Bregman et al.\ 1993; Hora et al.\ 1996).
Spectra taken at two different positions are shown, one centered on the
core, and the other at 4\arcsec\ S of the core. Both show strong warm
dust continuum, and the core also has strong CO bandhead emission features
in the $\lambda = 2.3 - 2.4$ \micron\ range.
\subsubsection{IRC+10$^{\circ}$420}
IRC+10$^{\circ}$420 is a highly evolved, OH/IR star that is thought to be in a
post-red supergiant phase (Jones et al.\ 1993). The central star seems to
have changed spectral type, transitioning recently to an early A type
(Oudmaijer 1998). Oudmaijer et al. (1996) detected several of the hydrogen
lines in absorption and emission in the near-IR, and Oudmaijer (1998)
presented a high-resolution 0.38 -- 1 \micron\ spectrum showing a large
number of emission and absorption lines. Recent HST imaging by Humphreys
et al.\ (1997) shows that the circumstellar environment around this star
is extremely complex, with spherical outer shells that extend to a
diameter of 6 arcseconds, and several inner condensations. In the near-
and mid-infrared, bipolar lobes are visible that extend $\sim 2$
arcseconds from the core.
The spectrum of IRC+10$^{\circ}$420 shown in Figure 31 was taken centered on the
object. The slit length includes the inner few arcseconds of the object,
although it is dominated by the bright core. The observed spectrum shows
a bright and relatively featureless continuum. Some \ion{H}{1} lines, e.g.,
Pa$\beta$, are seen in absorption.
\subsubsection{M 2--56}
The PPN M 2--56 is a bipolar nebula with a bright central core. It is
similar in morphology to AFGL 618, although it seems to be at an earlier
evolutionary stage since it does not appear to have an \ion{H}{2} region
(Trammell, Dinerstein, \& Goodrich 1993; Goodrich 1991). The spectrum of
M 2--56 shown in Figure 31 was taken centered on the core of this PPN. The
dominant feature is a hot dust continuum that is most prominent in the K
band region of the spectrum. There are some residual features in the
spectrum from imperfect sky subtraction, mostly in the K band.
\section{Discussion and Summary}
\subsection{Spectral Categories}
The PNe spectra presented in this paper were grouped according to spectral
characteristics as described above. The groups are an efficient way to
present the data, but also can be seen to correlate strongly with other
characteristics of the PN.
\subsubsection{Morphology}
The group of \ion{H}{1} - line dominated PNe is composed of primarily
elliptical or round PNe, along with the peculiar or irregular nebulae of
the sample. In general these PNe are well-known from optical studies,
identified either by their morphology or their optical spectra. Many of
the PNe in this group of the sample, however, do have IR ``excess''
continuum emission from warm dust, which in some cases prompted their
inclusion in this sample.
The spectral groups with molecular and/or dust continuum emission are
primarily bipolar. This classification includes objects such as NGC 6720
which have a ring morphology but are thought to be bipolar viewed pole-on;
Hb 12 which is brightest in H$_2$\ in the equatorial region and along the
outer edges of the lobes; M 2--9 which is brightest in H$_2$\ at the edges of
the lobes with no equatorial emission other than at the core; and AFGL
2688 which is brightest along the axis of the bipolar lobes, with H$_2$\
emission in the equatorial plane. Clearly this is a heterogeneous group
with a wide range of emission and morphological differences that imply a
range of evolutionary tracks and states.
\subsubsection{The Carbon-to-Oxygen Ratio}
Carbon stars, although a small fraction of all AGB stars, return about
half of the total mass injected into the ISM by all AGB stars, since
they have on average much higher mass loss rates ($>10^{-4}\ {\rm
M}_{\sun}\ {\rm yr}^{-1}$) than do O-rich objects. The
carbon-to-oxygen (C/O) abundance ratio in PNe has previously been
shown to correlate with morphology (Zuckerman \& Aller 1986), with
bipolar PNe tending to be carbon-rich. It is therefore expected that
the C/O ratio also correlates with the spectral classifications
presented here. This is in general the case, with the \ion{H}{1} -
line dominated PNe having C/O ratios less than or about 1, whereas the
remaining categories which are dominated by the bipolar PNe have C/O
ratios $>$ 1, as reported by Zuckerman \& Aller (1986) and Rola \&
Stasi\'nska (1994). Rola \& Stasi\'nska discuss problems with
previous determinations of the C/O ratio, and use different criteria
that result in a slightly lower percentage of carbon-rich PNe (35\%)
than others. Their ratios are used in the discussion below.
The morphology of PN also has been shown to depend on the progenitor mass
(see Corradi \& Schwarz 1995), with the bipolar PN being more massive than
other morphological types. This relationship, along with the link between
carbon abundance and morphology, suggests that carbon stars are the
progenitors of bipolar PN and those with a large amount of molecular
material. The mechanisms that cause massive carbon-rich stars to
preferentially form bipolar PN are still not understood.
There are some exceptions to the correlation of morphological type to C/O
ratio; in particular, NGC 6543 has a much higher value (9.55) than the
others in the class. In the other extreme, NGC 2346 stands out as having a
low C/O ratio (0.35) compared to other bipolar PNe in the H$_2$\ - dominated
group. This object is a much more evolved object than the others in its
group (e.g., AFGL 2688), and exhibits weak Pa$\beta$\ emission, showing that
the ionized gas is present although weak relative to the molecular
emission in the nebula.
\subsection{Spectral Sampling of Morphological Features}
This survey has differed from many previous investigations in that a
short, narrow slit was used to obtain the data, rather than a large beam
that could include most or all of the nebula. Because of this, one cannot
easily use the spectra presented here to model the PNe in a global sense,
if that requires a measurement of the total flux from the object. Also,
if a complete census of emission lines were required, some might be missed
if there were variations of emission characteristics across the nebula and
certain regions were not sampled.
The spatial selectivity that prevents viewing the entire PN at once,
however, has proven to be an advantage when trying to examine various
aspects of the PN, including variations across the nebula, as a function
of distance from the central star, or in examining certain morphological
features. For example, in M 2--9 and NGC 2440, the emission of the lobe
walls and emission knots were separately sampled, which showed the large
spectral differences in these regions. This information is important for
modeling the structure and formation of the PN. Another reason why the
small aperture is useful is that if the emission from a group of lines
such as H$_2$\ is to be modeled, it is important to compare the emission
from a clump of material where the conditions do not vary greatly over its
size. For example, the emission from H$_2$\ present very close to the
central star in a strong UV field could be quite different from H$_2$\
emission from the outer parts of the halo. Also, the small slit has aided
in detecting weak H$_2$\ emission from several PNe such as NGC 40, where
detection would have been difficult if the central star and the rest of
the nebula could not be excluded from the measurement.
\subsection{Summary of Molecular Hydrogen Emission in PNe}
A long-standing problem in the interpretation of H$_2$\ emission from
interstellar and circumstellar environments is understanding the
excitation mechanism. Three fundamental mechanisms are possible. One is
excitation of a near-infrared fluorescence spectrum resulting from a
rotational-vibrational cascade in the ground electronic state following
electronic excitation by the absorption a UV photon in the Lyman and
Werner bands (Black \& van Dishoeck 1987). A second excitation mechanism
is collisional excitation in a warm gas ($T_{\rm K} \gtrsim 1600$ K).
While UV excitation in a low density gas produces an easily identifiable
spectrum, the level populations can be driven to produce thermal line
ratios when the UV flux is large and densities begin to exceed $\approx$\
$10^4$ cm$^{-3}$ (Sternberg \& Dalgarno 1989). Detailed spectral and
morphological analysis are often required to determine an origin of the
near-IR spectrum. A third excitation mechanism is formation of H$_2$\ on
the surfaces of dust grains and in the gas phase. While potentially
important in isolated regions of certain objects, we do not consider this
to be generally important in PNe and PPNe relative to the other two
processes. This is because molecular formation in PNe is relatively slow
compared to dissociation rates.
In PNe and PPNe, the situation can be complicated by both dominant
excitation mechanisms being present simultaneously, and in different
forms. Several ways of exciting near-IR H$_2$\ emission have been
identified as possible: direct thermal excitation in warm gas created
behind moderate velocity shocks, direct excitation by UV photons from
the hot central star, somewhat indirectly by collisional excitation in
warm gas created by rapid grain streaming (e.g. Jura \& Kroto 1990), and
excitation through absorption of Ly$\alpha$~ photons (by an accidental
resonances with the B$^1\Sigma_u^+ - {\rm X}^1\Sigma_g^+$ $v = 1 - 2$ P(5)
and R(6) transitions of H$_2$) which can be generated in a nearby strong
shock (e.g.,\ Black \& van Dishoeck 1987).
The first two mechanisms have been identified in several PNe, such
as thermal excitation in AFGL 2688 (e.g.,\ Hora \& Latter 1994; Sahai et al.\
1998a), pure UV excitation in a low density gas around Hb 12 (e.g.,\
Dinerstein et al.\ 1988; Hora \& Latter 1996; Luhman \& Rieke 1996), and UV
excitation in a high density gas in M 2--9 (Hora \& Latter 1994) and NGC
7027 (Graham et al.\ 1993b; this paper). A combined spectrum was found from
a detailed analysis of AFGL 618 (Latter et al.\ 1992; this paper). While
the form of the excitation might be apparent for these and other objects, it
is not always evident what is the source of the warm gas or UV photons.
Winds are present in AFGL 2688 which could directly heat the gas through shocks,
but considerable grain streaming is likely taking place as well (see Jura
\& Kroto 1990).
Very fast winds and dissociating shocks are present in AFGL 618, M 2--9
(e.g.,\ Kelly, Latter, \& Hora 1998), and M 1--16 (Sahai et al.\ 1994; Schwarz
1992), and all show clear evidence of UV excitation. In addition, the
photon path to the H$_2$\ emitting regions is not in a direct
line-of-sight to the central star, which for photons coming from the
central star suggests scattering in what is a fairly low density medium.
Alternatively, we are seeing in each of these objects excitation of H$_2$\ at
the bipolar lobe walls by UV photons generated within strong shocks
produced by the fast winds. This hypothesis was explored through detailed
modeling by Latter et al.\ (1992) of AFGL 618, but the high relative
intensity of the thermally excited emission and poor spatial resolution
limited this analysis. The presence of very fast winds in the lobes of
each of these objects, and the presence of UV excited H$_2$\ emission at the
lobe walls strongly suggests that indirect excitation of the H$_2$\ is
occurring by interactions with photons generated by wind-produced
shocks. Detailed modeling of sensitive, high spatial resolution
spectra is required. It is also evident, in general, that without detailed
spectra, H$_2$\ is a rather poor diagnostic of overall conditions in PNe and
PPNe.
If we conclude that all of the ways to excite H$_2$\ in PNe and PPNe listed
above are present and important, what does this imply for our
understanding of these objects and the utility of H$_2$\ as a diagnostic? It
is now well understood that the presence of molecular emission from PNe
and PPNe is tied to the morphology of the objects such that if molecular
emission is present, the object has a bipolar morphology (e.g.,\ Zuckerman \&
Gatley 1988; Latter et al.\ 1996; Kastner et al.\ 1996, and references
therein). We have argued that H$_2$\ emission is excited in multiple
ways in PNe and PPNe. While special conditions are required for H$_2$
emission to be seen in near-IR spectra, the conditions that drive the
excitation are common in all PNe and PPNe and are not clearly dependent on
morphological type. A conclusion that can be drawn from this argument
alone is that molecular material is present in nebulae with a bipolar
morphology and a significant amount of molecular material is $not$ present
in other morphological types. Therefore, objects that have a bipolar
morphology must have a dense, high mass envelope in which the molecular
material can be shielded and survive dissociation for relatively long
times -- suggesting a high mass loss rate and a high mass progenitor star.
A correlation between bipolar morphology and high mass progenitor stars
has been found by others (e.g.,\ Corradi \& Schwarz 1995). It is apparent
that the presence of H$_2$\ emission in a PN is not tied to directly to the
morphology, but that the bipolar morphology is intimately related to the
density and mass of the circumstellar envelope, and therefore the mass of
the progenitor star. Why high mass, high mass loss rate asymptotic giant
branch stars shed material in an axisymmetric, not spherical, way remains
a mystery.
\acknowledgments
We thank Xander Tielens and David Hollenbach for useful discussions and
encouragement. We acknowledge support from NASA grant 399-20-61 from the
Long Term Space Astrophysics Program. WBL was supported during part of
this study by a National Research Council Research Associateship.
|
\section{Introduction.}
Soon after its early developments in probability theory [1] the study
of random walks and their applications has gradually carved its way into
the concerns of many topics in exact or applied science. In various
recent developments, the notion of ``walk'' in the proper sense of the
term has gradually blurred, giving way to a more general concept where
a ``step'' of the walk is not necessarily an algebraic distance between
two locations but may become a time interval between two consecutive
events of a chronological series, or even a sequence of real numbers.
Before mentioning a few significant fields of application, let us quote
more specifically the notion of ``correlated random walk,'' which underlies
the possibility for a step to depend on the values taken by some (or all) of
the previous steps. We will address the step correlations and their
consequences at length in the course of this paper, and will consider
therefore the ``direct'' problem, in contradistinction with the usual concern
prevailing in applied science which asks for answers to the ``inverse'' problem,
\textit{i.e.}, can a given set of data be suitably described by a (correlated or
uncorrelated) random walk formalism? Examples of fields resorting to the
inverse problem are neuromuscular [2] and cardiovascular medicine [3],
structural genetics [4-6], physiology [7], regulation of biological
rhythms [8]; to these fields pertaining essentially to the medical domain can be
added the soaring irruption of financial analysis [9-11] in view of market
predictions.
In this paper we will focus on the formalism of correlated random walks on
one-dimensional lattices and discuss the r\^ole of memory in some typical models
which can be solved exactly. With the help of a direct evaluation of the variance
of the walks, we will be able to prove that all walks with short-range memory
have a Hurst exponent equal to $\frac{1}{2}$ in the strict limit of an infinite
number of steps, although the apparent Hurst exponent may look different from
$\frac{1}{2}$. In particular, we will endeavor to produce a simple argument for
a necessary condition yielding a Hurst exponent larger than $\frac{1}{2}$.
\section{Correlated walks and memory}
The transition probabilities in a simple random walk on a
one-dimensional infinite lattice are constant and independent of the
course of the walk. If these transition probabilities depend upon the
$M$ previous steps, we have a correlated random walk whose memory range
is $M$. Correlated random walks were introduced by Taylor [12] in an analysis of
diffusion by continuous motions. The Taylor's model is a one-dimensional
random walk on a lattice, in which steps can be made to nearest neighbors
only. The walker has a probability $\alpha$ to repeat his previous move,
and a probability $1-\alpha$ to move in the opposite direction. The resulting
correlated random walk has a memory range equal to 1. If
$\alpha > \frac{1}{2}$ (resp. $\alpha < \frac{1}{2}$) the
walk is said to be {\it persistent} (resp. {\it antipersistent}). A complete
analysis of the Taylor's model has been given by Goldstein [13]. More examples
of persistent random walks have been studied by many other authors [14-16].
A detailed bibliography may be found in Weiss [17].
In this paper we study a class of correlated random walks with memory
range $M>1$. Our model may be described as follows. Let $(S_k)$ be
an infinite sequence of random variables such that, for $k>M$,
$$
S_k =
\begin{cases}
1 & \text{with probability $\frac{1}{2}p$}\\
-1 & \text{with probability $\frac{1}{2}p$}\\
f_M(S_{k-1}, S_{k-2},\ldots,S_{k-M}) & \text{with probability $q=1-p$,}\\
\end{cases}
$$
where $f_M$, called the {\it memory function}, is a Boolean function from
$\{-1,\!1\}^M$ onto $\{-1,\!1\}$.
For $k \leq M$, the $S_k$'s are independent, identically distributed (\textit{iid})
Bernoulli random variables, each
$S_k$ taking the values $\pm 1$ with equal probabilities. The values of these
$M$ random variables define the initial conditions for the walk. The position
of the random walker after $n$ steps is
$$
R_n = S_1 + S_2 + \cdots + S_n.
$$
The set $\mathcal{F}_M$ of all bounded functions defined on $\{-1,1\}^M$ is a
Euclidean space of dimension $2^M$ for the dot product defined by
\begin{eqnarray*}
\lefteqn{<f_M^{(1)}\mid f_M^{(2)}>\;\; = }\\
& & \left(\frac{1}{2}\right)^M
\tr f_M^{(1)}(S_{k-1}, S_{k-2},\ldots,S_{k-M})
f_M^{(2)}(S_{k-1}, S_{k-2},\ldots,S_{k-M}),\\
\end{eqnarray*}
where the trace operator $\tr$ is a sum over all the possible values of
the $M$ variables $S_{k-1}, S_{k-2},\ldots,S_{k-M}$.
The set of the following $2^M$ functions
$$\left\lbrace
\begin{array}{l}
1,\\
S_{k-1}, S_{k-2}, \ldots , S_{k-M},\\
S_{k-1}S_{k-2}, S_{k-1}S_{k-3}, \ldots , S_{k-M+1}S_{k-M},\\
S_{k-1}S_{k-2}S_{k-3}, \ldots , S_{k-M+2}S_{k-M+1}S_{k-M}, \\
\cdots,\\
S_{k-1}S_{k-2}\cdots S_{k-M},\\
\end{array}
\right.$$
is a complete orthonormal system, and any function in $\mathcal{F}_M$ can be
written as a linear combination of these $2^M$ functions. The most general
correlated random walk with a finite memory of range $M$ is, therefore,
defined by a sequence $(S_k)$ of random variables such that,
for $k>M$,
\begin{equation}
S_k =
\begin{cases}
1 & \text{with probability $\frac{1}{2}p$}\\
-1 & \text{with probability $\frac{1}{2}p$}\\
\varepsilon_{j_1j_2\cdots j_r}S_{k-j_1}S_{k-j_2}\cdots,S_{k-j_r}
& \text{with probability $qa_{j_1j_2\cdots j_r}$,}
\end{cases}
\label{general-model}
\end{equation}
where $p+q=1$, $1\leq j_1<j_2<\cdots<j_r\leq M$, and
$\varepsilon_{j_1j_2\cdots j_r}$, which are \textit{non-random}, are equal either to
$1$ or to $-1$. The $a$'s are conditional probabilities (some of them possibly
equal to zero) satisfying the completeness relation:
\begin{eqnarray*}
& & \sum_{1\leq j_1\leq M}a_{j_1} + \sum_{1\leq j_1<j_2\leq M}a_{j_1j_2}
+\cdots\\
& & \qquad + \sum_{1\leq j_1<j_2\leq<j_r\leq M} a_{j_1j_2\cdots j_r}
+ \cdots + a_{12\cdots M} = 1.\\
\end{eqnarray*}
As above, for $k \leq M$, $S_k$ is a simple Bernoulli random variable
which takes the values $\pm 1$ with equal probabilities.
In order to simplify this model, we require that, as in the case of
a simple symmetric random walk, the average value $<R_n>$ of the position
of the random walker after $n$ steps should be equal to zero. While this
condition is automatically satisfied if $n\leq M$, a tedious inspection
shows that, for $n>M$, the above requirement is satisfied if, and only if,
the memory function $f_M$ has the following simple forms:
\begin{itemize}
\item $f_M(S_{k-1}, S_{k-2},\ldots,S_{k-M})$ is a linear combination of
multilinear terms of odd degree in the $S_k$'s;
\item $f_M(S_{k-1}, S_{k-2},\ldots,S_{k-M})$ reduces to a single
multilinear term containing an even number of $S_k$'s.
\end{itemize}
For example, if $M=3$, the most general form for the memory function $f_3$,
satisfying the condition $<R_n>=0$, is either
\begin{eqnarray*}
f_3(S_{k-1},S_{k-2},S_{k-3})
& = & a_1\varepsilon_1S_{k-1}+a_2\varepsilon_2S_{k-2} + a_3\varepsilon_3S_{k-3}\\
& &\qquad + a_{123}\varepsilon_{123}S_{k-1}S_{k-2}S_{k-3},
\end{eqnarray*}
or
$$
f_3(S_{k-1},S_{k-2},S_{k-3}) = a_{ij}\varepsilon_{ij}S_{k-i}S_{k-j},
$$
with $1\leq i\leq 3$ and $1\leq j\leq 3$.
In the sequel we will only consider two cases: either $f_M$ is a linear
function of the $M$ previous steps, or the product of the $M$ previous steps.
Let us briefly examine these two cases.
In what follows $(X_k)$ denotes an infinite sequence of \textit{iid}
Bernoulli random variables such that, for all positive integers $k$, we have
$$
P(X_k=1) = P(X_k=-1) = \frac{1}{2},
$$
\textbf{1.} \textit{$f_M$ is linear.} If $(S_k)$ is an infinite sequence of
random variables defined by
\begin{equation}
S_k =
\begin{cases}
X_k & \text{if $k\le M$},\\[0.1cm]
X_k & \text{with probability $p$, if $k>M$},\\
\varepsilon_j S_{k-j} & \text{with probability $qa_j$, if $k>M$,}
\end{cases}
\label{general-linear}
\end{equation}
where, for $j=1,2,\ldots,M$, the $\varepsilon_j$ have fixed values chosen
\textit{a priori\/} equal either to $1$ or $-1$, and the $a_j$ are $M$
nonnegative real numbers such that
$$
a_1+a_2+\cdots+a_M = 1.
$$
It is clear that, for $k\leq M$, $<S_k>=0$. Hence
\begin{eqnarray*}
<S_{M+1}> & = & p <X_{M+1}>+qa_1\varepsilon_1<S_M>
+\cdots+qa_M\varepsilon_M<S_1>\\
& = & 0\\
<S_{M+2}> & = & p <X_{M+2}>+qa_1\varepsilon_1<S_{M+1}>
+\cdots+qa_M\varepsilon_M<S_2>\\
& = & 0\\
& & \cdots\\
<S_{M+\ell}> & = & p <X_{M+\ell}>+qa_1\varepsilon_1<S_{M+\ell-1}>
+\cdots+qa_M\varepsilon_M <S_\ell>\\
& = & 0,
\end{eqnarray*}
and, therefore, $<R_n>=0$, for all positive integers $n$.
\textbf{2.} \textit{$f_M$ is a single product of $r$ terms.} If $(S_k)$ is
an infinite sequence of random variables defined by
\begin{equation}
S_k =
\begin{cases}
X_k & \text{if $k\le M$},\\[0.1cm]
X_k & \text{with probability $p$ if $k>M$},\\
\varepsilon_{j_1j_2\cdots j_r}S_{k-j_1}S_{k-j_2}\cdots S_{k-jr}
& \text{with probability $q$ if $k>M$},
\end{cases}
\label{general-product}
\end{equation}
where $q=1-p$, $1\leq j_1<j_2<\cdots<j_r\leq M$, and
$\varepsilon_{j_1j_2\cdots j_r}$ is given and equal to either $1$ or $-1$.
Here again, for $k\leq M$, $<S_k>=0$. Hence, assuming first that $r=M$,
we have
\begin{eqnarray*}
<S_{M+1}>& = & p<X_{M+1}>+q\varepsilon_{12\cdots M}<S_1S_2\cdots S_M>\\
& = & p<X_{M+1}>+q\varepsilon_{12\cdots M}<S_1><S_2>\cdots <S_M>\\
& = & 0\\
<S_{M+2}>& = & p<X_{M+2}>+q\varepsilon_{12\cdots M}
<S_2S_3\cdots S_M S_{M+1}>\\
& = & pq\varepsilon_{12\cdots M}<S_2S_3\cdots S_M X_{M+1}>\\
& & \qquad + q^2\varepsilon_{12\cdots M}<S_1S_2^2S_3^2\cdots S_M^2>\\
& = & q^2\varepsilon_{12\cdots M}<S_1>\;=\;0
\end{eqnarray*}
where in the expression of $S_{M+2}$ we have replaced $S_{M+1}$ by its
expression. More generally,
\begin{eqnarray*}
<S_{M+\ell}> & = & p<X_{M+\ell}>+q<S_\ell S_{\ell+1}\cdots S_{M+\ell-1}>\\
& = & pq<S_\ell S_{\ell+1}\cdots X_{M+\ell-1}>+
q^2<S_{\ell-1}S_\ell^2 S_{\ell+1}^2\cdots S_{M+\ell-2}^2>\\
& = & q^2<S_{\ell-1}>,
\end{eqnarray*}
which shows that for all positive integers $k$, $<S_k>=0$,
and, consequently, $<R_n>=0$.
Along similar lines, it can be shown that the result $<R_n>=0$ remains
valid for $r<M$, although its proof requires a larger number of substitutions.
In the following sections we will show that, for any finite value of $M$,
we can determine exactly the probability distribution of $<R_n>$,
using a transfer matrix method. It will turn out that the rank of the matrix
is equal to $2^M$. Due to the increasing complexity of the eigensystem problem
with increasing $M$, this method becomes rapidly cumbersome, although it
rather easily provides a numerical solution if the number of steps $n$ is
not too large. In order to illustrate the method, we will first briefly review
the Goldstein's solution of the Taylor's model.
Since the asymptotic behavior of the variance $<R_n^2>$ is of particular
interest to natural scientists,
we will directly derive its exact expression through a recursive method
that applies at any finite space dimensionality. From this expression we
will be able to show that the asymptotic behavior $<R_n^2>$,
when $n$ goes to infinity, exhibits a crossover phenomenon.
\section{Taylor's model.}
The Taylor's model corresponds to Model~\ref{general-linear} for $M=1$,
that is,
$$
S_1 = X_1,
$$
and, for all $k>1$,
$$
S_k=
\begin{cases}
X_k, & \text{with probability $p$}\\
\varepsilon S_{k-1}, & \text{with probability $q=1-p$,}
\end{cases}
$$
where $(X_k)$ is a sequence of \textit{iid} Bernoulli variables such that,
for all positive integers $k$,
$$
P(X_k=-1) = P(X_k=1) = \frac{1}{2},
$$
and $\varepsilon$ is a \textit{non-random number} equal to $\pm 1$.
If $p_n(k)$ denotes the probability that the walker will be at site $k$ after
the $n$th step has been completed, we have
$$
p_n(k) = p_n^+(k)+p_n^-(k),
$$
where the superscript $\pm$ refers to the sign of the $n$-th step, that is
$$
p^{\pm}(k,n) = P((R_n=k)\cap (S_n = \pm 1)).
$$
Taking into account the conditional probabilities:
\begin{eqnarray*}
P(S_n=1\mid S_{n-1}=1) = P(S_n=-1\mid S_{n-1}=-1) & = &
\frac{1}{2}\left(p+(1+\varepsilon)q\right)\\
P(S_n=1\mid S_{n-1}=-1) = P(S_n=-1\mid S_{n-1}=1) & = &
\frac{1}{2}\left(p+(1-\varepsilon)q\right),
\end{eqnarray*}
we obtain the following recursion relations
\begin{eqnarray}
p_n^+(k) & = & \frac{1}{2}\left(p+(1+\varepsilon)q\right)p_{n-1}^+(k-1)+
\frac{1}{2}\left(p+(1-\varepsilon)q\right)p_{n-1}^-(k-1)\nonumber\\
& = & \frac{1}{2}\left(1+\varepsilon q\right)p_{n-1}^+(k-1)+
\frac{1}{2}\left(1-\varepsilon q\right)p_{n-1}^-(k-1)\label{pplus}\\
p_n^-(k) & = & \frac{1}{2}\left(p+(1-\varepsilon)q\right)p_{n-1}^+(k+1)+
\frac{1}{2}\left(p+(1+\varepsilon)q\right)p_{n-1}^-(k+1)\nonumber\\
& = & \frac{1}{2}\left(1-\varepsilon q\right)p_{n-1}^+(k+1)+
\frac{1}{2}\left(1+\varepsilon q\right)p_{n-1}^-(k+1)\label{pminus},
\end{eqnarray}
or, in a more condensed form which will be useful when we consider
the case $M>1$,
\begin{equation}
p_n^\sigma(k)= \frac{1}{2}\left(1-\varepsilon q\right)p_{n-1}^+(k-\sigma)+
\frac{1}{2}\left(1+\varepsilon q\right)p_{n-1}^-(k-\sigma),
\label{psigma}
\end{equation}
with $\sigma=\pm1$.
The probability $\alpha$ in Goldstein's notations coincides here with
$ \frac{1}{2}\left(1+\varepsilon q\right)$.
The generating function of the probability distribution of the
random walk is defined by
$$
f_n(x) = f_n^+(x) + f_n^-(x),
$$
with
\begin{eqnarray*}
f_n^+(x) & = & \sum_k p_n^+(k)\; x^{k}\\
f_n^-(x) & = & \sum_k p_n^-(k)\; x^{k},
\end{eqnarray*}
where the summation index $k$ runs from $-n$ to $n$ with step 2.
Let
$$
\mathbf{f}_n(x) =
\left(\begin{array}{c}
f_n^+(x)\\[0.2cm]
f_n^-(x)
\end{array}\right)
$$
and
$$
{\mathbf M}(x) =
\left(\begin{array}{cc}
\frac{1}{2}\left(1+\varepsilon q\right)\; x
& \frac{1}{2}\left(1-\varepsilon q\right) \; x\\
\\
\frac{1}{2}\left(1-\varepsilon q\right) \; \frac{1}{x}
& \frac{1}{2}\left(1+\varepsilon q\right)\; \frac{1}{x}\\
\end{array}\right).
$$
Then
\begin{equation}
\mathbf{f}_n(x) = \mathbf{M}(x) \; \mathbf{f}_{n-1}(x).
\label{1fiter}
\end{equation}
Iterating this recursion relation, we obtain
\begin{equation}
\mathbf{f}_n(x) = \mathbf{M}^{n-1}(x) \; \mathbf{f}_1(x),
\label{recur-1fn}
\end{equation}
where
$$
\mathbf{f}_1(x) =
\left(\begin{array}{c}
f_{1}^+(x)\\[0.2cm]
f_{1}^-(x)
\end{array}\right)=
\left(\begin{array}{c}
\frac{x}{2}\\[0.2cm]
\frac{1}{2x}
\end{array}\right).
$$
Let $\mathbf{R}(x)$ be the transformation matrix such that
$$
\mathbf{D}(x) = \mathbf{R}^{-1}(x) \; \mathbf{M}(x) \; \mathbf{R}(x)
=\left(\begin{array}{cc}\lambda_1(x) & 0\\[0.2cm]
0 & \lambda_2(x)
\end{array}\right),
$$
where $\lambda_1(x)$ and $\lambda_2(x)$ are the eigenvalues of $\mathbf{M}(x)$.
Then
$$
\mathbf{f}_n(x) =
\mathbf{R}(x)\; \mathbf{D}^{n-1}(x)\; \mathbf{R}^{-1}(x)\; \mathbf{f}_1(x)
$$
The above relation involves the eigenvalues and eigenvectors of
$\mathbf{M}(x)$ which can be readily obtained.
However, the exact expression
for the generating function of $p_n(k)$ is cumbersome to manipulate,
and does not easily yield the variance. Therefore, we resort to an
alternative and more direct method, which allows for a recursive
evaluation of $<R_n^2>$:
\begin{eqnarray*}
<R_n^2> & = & <(S_1+S_2+\cdots+S_n)^2>\\
& = & <(S_1+S_2+\cdots+S_{n-1})^2> \\
& & \qquad + 2<S_n\;(S_1+S_2+\cdots+S_{n-1})> + <S_n^2>\\
& = & <R_{n-1}^2> + 2<S_n\;(S_1+S_2+\cdots+S_{n-1})> + <S_n^2>.\\
\end{eqnarray*}
As will appear below, the correlation coefficient $<S_n S_{n-k}>$ does not
depend explicitly upon $n$, thus we use the notation
$$
c_k = <S_n S_{n-k}>,
$$
and obtain
$$
<R_n^2> = <R_{n-1}^2> + 2 (c_1+c_2 + \cdots + c_{n-1}) + 1,
$$
where we have taken into account that $<S_n^2>=1$. Evaluating the $c_k$'s,
we get
\begin{eqnarray*}
c_1 & = & <S_nS_{n-1}> = p\;<X_nS_{n-1}> + \;\varepsilon q\;<S_{n-1}^2>
=\varepsilon q\\
c_2 & = & <S_nS_{n-2}> = p\;<X_nS_{n-2}> + \; \varepsilon q\;<S_{n-1}S_{n-2}>
=\varepsilon qc_1=q^2\\
& \cdots & \\
c_k & = & <S_nS_{n-k}> = p\;<X_nS_{n-k}> + \varepsilon q\;<S_{n-1}S_{n-k}>
=\varepsilon qc_{k-1}=(\varepsilon q)^k.
\end{eqnarray*}
Substituting the $c_k$'s in the expression of $<R_n^2>$ we obtain
\begin{eqnarray*}
<R_n^2> & = & <R_{n-1}^2> +\; 2\;(\varepsilon q+q^2+\cdots
+(\varepsilon q)^{n-1}) + 1\\
& = & <R_{n-1}^2> +
\; 2\;\frac{\varepsilon q-(\varepsilon q)^n}{1-\varepsilon q}+1,\\
\end{eqnarray*}
and iteration of this recursion relation yields
\begin{eqnarray}
<R_n^2> & = & n + \frac{2}{(1-\varepsilon q)^2}
\;\left((n-1)\varepsilon q - nq^2+(\varepsilon q)^{n+1}\right)\nonumber\\
& = & n + \frac{2}{(1-\varepsilon (1-p))^2}\times\nonumber\\
& & \quad \left((n-1)\varepsilon (1-p) - n(1-p)^2
+(\varepsilon)^{n+1}(1-p)^{n+1}\right).\label{varofp}
\end{eqnarray}
It is quite remarkable that for a similar random walk on a $d$-dimensional
simple cubic lattice, the above result is preserved. Of course, in this case,
the sequence of one-dimensional $X_k$'s is replaced by a sequence of
$d$-dimensional random vectors ${\mathbf X}_k$'s, and each elementary step
can be performed along each of the $d$ axes, either in the positive or the
negative direction.
For any fixed nonzero value of $p$, when $n$ goes to infinity, the term
$(1-p)^{n+1}$ goes to zero, and $<R^2_n>$ behaves as
$$
\frac{1+\varepsilon q}{1-\varepsilon q}\;n,
$$
and the random walk is eventually Gaussian at very large $n$.
However, when $q$ goes to 1, according to whether $\varepsilon$ is equal to
$+1$ or $-1$, the prefactor of $n$ goes, respectively, either to infinity
or to zero. Consequently, if $p$ is very small, the asymptotic behavior
of $<R_n^2>$ should be carefully analyzed, distinguishing the two cases
$\varepsilon=1$ and $\varepsilon=-1$.
\subsubsection{Persistent random walk ($\varepsilon=1$).} Replacing $\varepsilon$ by $+1$ in
\ref{varofp}, and expanding the resulting expression in powers of $p$,
we obtain
$$
<R_n^2> = n + \frac{2}{p^2}\left(\frac{n^2p^2}{2}-\frac{np^2}{2}
-\frac{n(n^2-1)p^3}{6}+O(p^4)\right),
$$
which shows that, as expected, when $p$ tends to zero, $<R_n^2>$ tends to
$n^2$. Therefore, for a small fixed value of $p$, when $n$ increases,
we expect, for the variance, a crossover from a $n^2$ behavior to an $n$
behavior. This crossover is controlled by the term $(1-p)^{n+1}$ in
\ref{varofp}, and the above expansion of $<R_n^2>$ shows that the relevant
parameter is the variable $a=n p$. To characterize this behavior, let
define the exponent
$$
E(n,p) = \frac{\partial\log<R_n^2>}{\partial\log n},
$$
which is twice the Hurst exponent. A simple calculation yields
$$
\lim_{p\to 0} E\left(\frac{a}{p}, p\right)
= \frac{a(1-\mathrm{e}^{-a})}{a+\mathrm{e}^{-a}-1}.
$$
This expression shows that, if $n$ is large but small compared to $1/p$ (\textit{i.e.},
$a$ is small), the exponent $E$ tends to 2, whereas if $n$ is large compared to
$1/p$ ($a$ is large), $E$ tends to 1. This result is illustrated in
Figure~\ref{fig:M1H5m5+} for $p=0.00005$. Note that the crossover takes
place around $n=\mathrm{e}^{11}\approx 6\,10^4$, $np$ being of the order of unity.
\begin{figure}[h]
\centerline{\includegraphics*[scale=0.8]{M1H5m5+.eps}}
\caption{\label{fig:M1H5m5+}\textit{Exponent $E(n,p)$ of the persistent walk
as a function of $\log n$, for $p=5\,10^{-5}$ and $n_{\max}=1.6\,10^7$.}}
\end{figure}
The crossover from $E(n,p)=2$ to $E(n,p)=1$, clearly observed in the above
figure, where $p$ has a small fixed value and $n$ varies, can also be evidenced
from the figures representing
the probability distribution for a fixed value of $p$:
\begin{itemize}
\item Figure~~\ref{fig:M1n100p2m3+}: at small $p$ ($np\ll 1$) most of the
weight of the distribution is concentrated at the edges of the distribution range.
\item Figure~\ref{fig:M1n100p2m1+}: at large $p$ ($np\gg 1$) the significant
weight of the distribution spreads out in a Gaussian-like profile, with a
width roughly proportional to $\sqrt{n}$.
\end{itemize}
\vspace{0.3cm}
\begin{figure}[h]
\noindent
\begin{minipage}{0.46\linewidth}
\centering\epsfig{figure=M1n100p2m3+.eps, width=\linewidth}
\caption{\label{fig:M1n100p2m3+}\textit{$p_n(k)$ for $M=1$,
$\varepsilon=1$, $p=0.002$ and $n=100$.}}
\end{minipage}\hfill
\begin{minipage}{0.46\linewidth}
\centering\epsfig{figure=M1n100p2m1+.eps, width=\linewidth}
\caption{\label{fig:M1n100p2m1+}\textit{$p_n(k)$ for $M=1$,
$\varepsilon=1$, $p=0.2$ and $n=100$.}}
\end{minipage}
\end{figure}
The fulfillment of both conditions (with perhaps the possibility that, in
the case corresponding to $np\ll 1$, the localization takes place at various
points of the distribution range, and not necessarily at the edges,
as, for example, for the antipersistent walk discussed below) will be from
now on considered as a sufficient condition for the occurrence of a crossover.
\subsubsection{Antipersistent random walk ($\varepsilon=-1$).} Clearly, if
$p=0$ the walk will be a zigzag
walk, and $<R_n^2>=0$ (resp. $<R_n^2>=1$) if $n$ is even (resp. odd). If $p$
is small, a calculation similar to the previous one shows that:
$\bullet$ If $n p\gg 1$, $<R_n^2>$ behaves as $\frac{1}{2}n p$, and the
oscillatory terms related to the parity of $n$ are negligible.
$\bullet$ If $n p\ll 1$, we find that $<R_n^2>$ behaves as $n p$ when $n$ is
even, and as $1$ when $n$ is odd. Both expressions are the order of unity or
less, so that $<R_n^2>$ behaves as $n^0$.
Figure\ref{fig:M1H5m5-} represents the exponent characterizing the behavior
of the variance. In order to avoid, for values of $n$ such that $np\ll 1$,
the spurious behavior of $<R_n^2>$ in $n$ for even values of $n$, we have
only considered the odd values of $n$.
\begin{figure}[h]
\centerline{\includegraphics*[scale=0.8]{M1H5m5-.eps}}
\caption{\label{fig:M1H5m5-}\textit{Exponent $E(n,p)$ of the antipersistent walk
as a function of $\log n$ ($n$ odd),
for $p=5\,10^{-5}$ and $n_{\max}=1.6\,10^7$.}}
\end{figure}
\section{Linear memory function.}
We first address explicitly the case $M=2$, as a significant illustration
of the complexity which manifests as soon as $M>1$. Even in this case, a closed
form for the probability distribution is already difficult to write
down. However, we provide a formalism that can be fruitfully applied, at least
for numerical purposes, to larger values of $M$.
\subsection{Probability distribution for $M=2$.}
Let $p^{\sigma_1\sigma_2}(k,n)$ denote the probability that, after
$n$ steps, the random walker is at site $k$ and the steps $S_{n-1}$
and $S_n$ are , respectively, equal to $\sigma_1$ and $\sigma_2$, that is
$$
p_n^{\sigma_1\sigma_2}(k)=P((R_n=k)\cap ((S_{n-1}=\sigma_1)\cap(S_n=\sigma_2))).
$$
If $S_{n-1}=\sigma_1$ and $S_n=\sigma_2$, then $R_{n-1}=k-\sigma_2$, and
$p_n^{\sigma_1\sigma_2}(k)$ is a linear combination of two probabilities
of the form $p_{n-1}^{\sigma\sigma_1}(k-\sigma_2)$ where $\sigma=\pm 1$:
\begin{eqnarray}
p_n^{\sigma_1\sigma_2}(k) & = & \frac{1}{2}
\left(1+q(a_1\sigma_1\sigma_2\varepsilon_1+a_2\sigma_2\varepsilon_2)\right)
p_{n-1}^{+\sigma_1}(k-\sigma_2)\nonumber\\
& & \quad + \frac{1}{2}
\left(1+q(a_1\sigma_1\sigma_2\varepsilon_1-a_2\sigma_2\varepsilon_2)\right)
p_{n-1}^{-\sigma_1}(k-\sigma_2).
\label{p12}
\end{eqnarray}
To this equation we associate the equation expressing
$p_n^{\sigma_1{\overline{\sigma_2}}}(k)$, where ${\overline\sigma_2}=-\sigma_2$,
since this probability is a linear combination of the two probabilities
$p_{n-1}^{\sigma\sigma_1}(k-\overline{\sigma_2}) =
p_{n-1}^{\sigma\sigma_1}(k+\sigma_2)$ where $\sigma=1$ or $-1$.
We have
\begin{eqnarray}
p_n^{\sigma_1{\overline\sigma_2}}(k) & = & \frac{1}{2}
\left(1+q(a_1\sigma_1{\overline\sigma_2}\varepsilon_1
+a_2{\overline\sigma_2}\varepsilon_2)\right)
p_{n-1}^{+\sigma_1}(k-{\overline\sigma_2})\nonumber\\
& & \quad + \frac{1}{2}
\left(1+q(a_1\sigma_1{\overline\sigma_2}\varepsilon_1
-a_2{\overline\sigma_2}\varepsilon_2)\right)
p_{n-1}^{-\sigma_1}(k-{\overline\sigma_2}).
\label{p12bar}
\end{eqnarray}
The generating function of the probability distribution of $<R_n>$ is defined by
$$
f_n(x) = f_n^{++}(x) + f_n^{+-}(x) + f_n^{-+}(x) + f_n^{--}(x),
$$
with
$$
f_n^{\sigma_1\sigma_2}(x) = \sum_k p_n^{\sigma_1\sigma_2}(k)\; x^{k},
$$
where the summation index $k$ runs from $-n$ to $n$ with step 2.
From Equations~\ref{p12} and \ref{p12bar}, replacing $\sigma_1$ and
$\sigma_2$ by all their possible respective values, we obtain
\begin{equation}
\left(\begin{array}{c}
f_n^{++}(x)\\[0.2cm]
f_n^{+-}(x)
\end{array}\right)
=
\mathbf{M^+}
\left(\begin{array}{c}
f_{n-1}^{++}(x)\\[0.2cm]
f_{n-1}^{-+}(x)
\end{array}\right),
\label{fMplus}
\end{equation}
where
\begin{equation}
\mathbf{M^+}(x) = \left(\begin{array}{cc}
\frac{1}{2}
\left(1+q(a_1\varepsilon_1+a_2\varepsilon_2)\right)\; x
& \frac{1}{2}
\left(1+q(a_1\varepsilon_1-a_2\varepsilon_2)\right)\;x\\[0.2cm]
\frac{1}{2}
\left(1-q(a_1\varepsilon_1
+a_2\varepsilon_2)\right)\; \frac{1}{x}
&\frac{1}{2}
\left(1-q(a_1\varepsilon_1
-a_2\varepsilon_2)\right)\;\frac{1}{x}
\end{array}\right),
\label{Mplus}
\end{equation}
and
\begin{equation}
\left(\begin{array}{c}
f_n^{-+}(x)\\[0.2cm]
f_n^{--}(x)
\end{array}\right)
=
\mathbf{M^-}
\left(\begin{array}{c}
f_{n-1}^{+-}(x)\\[0.2cm]
f_{n-1}^{-+}(x)
\end{array}\right),
\label{fMminus}
\end{equation}
where
\begin{equation}
\mathbf{M^-}(x) = \left(\begin{array}{cc}
\frac{1}{2}\left(1-q(a_1\varepsilon_1-a_2\varepsilon_2)\right)\; x
&\frac{1}{2}\left(1-q(a_1\varepsilon_1+a_2\varepsilon_2)\right)\; x\\[0.2cm]
\frac{1}{2}\left(1+q(a_1\varepsilon_1-a_2\varepsilon_2)\right)\;\frac{1}{x}
&\frac{1}{2}\left(1+q(a_1\varepsilon_1+a_2\varepsilon_2)\right)
\;\frac{1}{x}
\end{array}\right).
\label{Mminus}
\end{equation}
Equations~\ref{fMplus} and \ref{fMminus} can be grouped together in a unique
equation that reads
$$
\left(\begin{array}{c}
f_n^{++}(x)\\[0.18cm]
f_n^{+-}(x)\\[0.18cm]
f_n^{-+}(x)\\[0.18cm]
f_n^{--}(x)
\end{array}\right)
=
\left(\begin{array}{cc}
\mbox{\huge{$\mathbf{M^+}(x)$}} & \mbox{\huge{$\mathbf{0}$}}\\[0.8cm]
\mbox{\huge{$\mathbf{0}$}} & \mbox{\huge{$\mathbf{M^-}(x)$}}
\end{array}\right)
\left(\begin{array}{c}
f_{n-1}^{++}(x)\\[0.18cm]
f_{n-1}^{-+}(x)\\[0.18cm]
f_{n-1}^{+-}(x)\\[0.18cm]
f_{n-1}^{--}(x)
\end{array}\right).
$$
Since the column vector on the right hand side of the above equation is not
correctly ordered, we have to introduce a $4\times 4$ permutation matrix, and
write this equation under the final form
\begin{equation}
\mathbf{f}_n(x) = \mathbf{M}(x)\;\mathcal{P}_4\; \mathbf{f}_{n-1}(x)
\label{2fiter}
\end{equation}
where
$$
\mathbf{f}_n(x) =
\left(\begin{array}{c}
f_n^{++}(x)\\[0.2cm]
f_n^{+-}(x)\\[0.2cm]
f_n^{-+}(x)\\[0.2cm]
f_n^{--}(x)\\
\end{array}\right),
$$
$$
\mathbf{M}(x) =
\left(\begin{array}{cc}
\mathbf{M^+}(x) & \mathbf{0}\\[0.2cm]
\mathbf{0} & \mathbf {M^-}(x)
\end{array}\right)
$$
and
$$
\mathcal{P}_4=
\left(\begin{array}{cccc}
1 & 0 & 0 & 0\\
0 & 0 & 1 & 0\\
0 & 1 & 0 & 0\\
0 & 0 & 0 & 1
\end{array}\right).
$$
Iterating Relation~\ref{2fiter}, we obtain
\begin{equation}
\mathbf{f}_n(x)
= \left(\mathbf{M}(x)\;\mathcal{P}_4\right)^{n-2} \; \mathbf{f}_2(x),
\label{recur-2fn}
\end{equation}
where
$$
\mathbf{f}_2(x) =
\left(\begin{array}{c}
f_{2}^{++}(x)\\[0.2cm]
f_{2}^{+-}(x)\\[0.2cm]
f_{2}^{-+}(x)\\[0.2cm]
f_{2}^{--}(x)\\
\end{array}\right)=
\left(\begin{array}{c}
\frac{x^2}{4}\\[0.2cm]
\frac{1}{4}\\[0.2cm]
\frac{1}{4}\\[0.2cm]
\frac{1}{4x^2}
\end{array}\right).
$$
\vspace{0.3cm}
\begin{figure}[t]
\noindent
\begin{minipage}{0.46\linewidth}
\centering\epsfig{figure=LM2n50p5m3++.eps, width=\linewidth}
\caption{\label{fig:LM2n50p5m3++}\textit{$p_n(k)$ for
$M=2$, $a_1=0.5$, $a_2=0.5$, $\varepsilon_1=1$, $\varepsilon_2=1$, $p=0.005$,
and $n=50$.}}
\end{minipage}\hfill
\begin{minipage}{0.46\linewidth}
\centering\epsfig{figure=LM2n50p5m1++.eps, width=\linewidth}
\caption{\label{fig:LM2n50p5m1++}\textit{$p_n(k)$ for
$M=2$, $a_1=0.5$, $a_2=0.5$, $\varepsilon_1=1$, $\varepsilon_2=1$, $p=0.5$,
and $n=50$.}}
\end{minipage}
\end{figure}
Further progress toward the expression of the generating function $f_n(x)$
requires the explicit diagonalization of the $4\times 4$ matrix
${\mathbf M}(x)\mathcal{P}_4$, followed by the procedure outlined in the $M=1$
case. The formal derivation of $f_n(x)$ is extremely complicated, and would
reveal untractable for $M>2$ as will be seen below.
However, the formalism is well suited to numerically evaluate the probability
distribution of $R_n$ recursively using~\ref{recur-2fn}.
Figures~\ref{fig:LM2n50p5m3++} and \ref{fig:LM2n50p5m1++}
show the probability distribution $p_n(k)$ when $a_1=a_2=0.5$ and
$\varepsilon_1=\varepsilon_2=1$
for two values of $p$ to illustrate cases $np\ll 1$ and $np\gg 1$.
Clearly, at the light of the comment at the end of subsection 3.0.1, the set
of these two figures is a sure sign of the existence of a crossover.
\subsection{Variance for $M=2$.}
Correlations are more tedious to evaluate than in the case $M=1$. Let
$$
c_{1,\ell} = <S_\ell S_{\ell+1}>;
$$
then
\begin{eqnarray*}
c_{1,1} & = & <S_1 S_2> = 0\\
c_{1,2} & = & <S_2 S_3> = <S_2 (p X_3 +
q(a_1\varepsilon_1S_2 + a_2\varepsilon_2 S_1))> = qa_1\varepsilon_1
\end{eqnarray*}
and, for $\ell\ge 3$,
\begin{eqnarray*}
c_{1,\ell} & = & <S_\ell S_{\ell+1}>
= <S_\ell (pX_{\ell+1} + qa_1\varepsilon_1 S_\ell
+ qa_2\varepsilon_2 S_{\ell-1})> \\
& = & qa_1\varepsilon_1 + qa_2\varepsilon_2 c_{1,\ell-1}.
\end{eqnarray*}
Hence
\begin{equation}
c_{1,\ell} = qa_1\varepsilon_1
\frac{1-(qa_2\varepsilon_2)^{\ell-1}}{1-qa_2\varepsilon_2}\quad(\ell\ge 1)
\label{correl-1ell}
\end{equation}
Similarly, if
$$
c_{2,\ell} = <S_\ell S_{\ell+2}>;
$$
then, for all positive integers $\ell$,
\begin{eqnarray*}
c_{2,\ell} & = & <S_\ell S_{\ell+2}>
= <S_\ell (pX_{\ell+2} + q(a_1\varepsilon_1 S_{\ell+1}
+ a_2\varepsilon_2 S_\ell))>\\
& = & qa_1\varepsilon_1 c_{1,\ell} + qa_2\varepsilon_2
\end{eqnarray*}
so that
\begin{eqnarray}
c_{2,\ell} = \frac{(qa_1\varepsilon_1)^2}
{1-qa_2\varepsilon_2}((1-qa_2\varepsilon_2)^{\ell-1}) + qa_2\varepsilon_2
\quad(\ell\ge 1).
\label{correl-2ell}
\end{eqnarray}
More generally, for $r>2$,
\begin{eqnarray}
c_{r,\ell} & = & <S_\ell S_{\ell+r}> = <S_\ell (p X_{\ell+r} +
q(a_1\varepsilon_1 S_{k+r-1} + a_2\varepsilon_2 S_{k+r-2}))>\nonumber\\
& = & q(a_1\varepsilon_1 c_{r-1,\ell} + a_2\varepsilon_2 c_{r-2,\ell}).
\label{correl-rell}
\end{eqnarray}
From this recursion relation, we obtain the following expression of $c_{r,\ell}$
\begin{equation}
c_{r,\ell} = \frac{c_{1,\ell}\lambda_2-c_{2,\ell}}{\lambda_2-\lambda_1}
\lambda_1^{r-1}
+ \frac{c_{2,\ell}-c_{1,\ell}\lambda_1}{\lambda_2-\lambda_1}
\lambda_2^{r-1},\label{correlofr}
\end{equation}
where $\lambda_1$ and $\lambda_2$ are the roots of
$$
\lambda^2-qa_1\varepsilon_1\lambda-qa_2\varepsilon_2=0.
$$
It is easy to verify that $|\lambda_1|\le 1$ and $|\lambda_2|\le 1$, the equal
sign being possible if $p=0$. For $p>0$, these conditions imply that the
correlations $c_{r,\ell}$ decrease exponentially in $r$, and this, in turn,
will ensure a linear dependence in $n$ for the variance.
Since the variance is given by
\begin{eqnarray*}
<R_n^2> & = & <(S_1+S_2+\cdots+S_n)^2>\nonumber\\
& = & n + 2\sum_{\ell=1}^{n-1}\sum_{r=1}^{n-\ell}c_{r,\ell},
\label{LMV}
\end{eqnarray*}
to find its its explicit expression, we need to evaluate the double summation.
Summing first over $r$, we obtain
\begin{small}
\begin{eqnarray*}
\sum_{r=1}^{n-\ell}c_{r,\ell} & = &
\frac{1-\lambda_1^{n-\ell}}{(\lambda_2-\lambda_1)(1-\lambda_1)}
\left(\frac{qa_1\varepsilon_1}{1-qa_2\varepsilon_2}
(1-(qa_2\varepsilon_2)^{\ell-1})(\lambda_2-qa_1\varepsilon_1)-qa_2\varepsilon_2\right)\\
& & + \frac{1-\lambda_2^{n-\ell}}{(\lambda_2-\lambda_1)(1-\lambda_2)}
\left(\frac{qa_1\varepsilon_1}{1-qa_2\varepsilon_2}
(1-(qa_2\varepsilon_2)^{\ell-1})(qa_1\varepsilon_1-\lambda_1)+qa_2\varepsilon_2\right),
\end{eqnarray*}
\end{small}
where we have replaced $c_{2,\ell}$ by its expression given by \ref{correl-2ell}.
Summing now over $\ell$ finally yields
\begin{small}
\begin{eqnarray*}
<R_n^2> & = & n + \frac{2(n-1)}{(\lambda_2-\lambda_1)(1-\lambda_1)}\left(
\frac{qa_1\varepsilon_1(\lambda_2-qa_1\varepsilon_1)}{1-qa_2\varepsilon_2}-qa_2\varepsilon_2
\right)\nonumber\\
& & - \frac{2}{(\lambda_2-\lambda_1)(1-\lambda_1)}\times
\frac{qa_1\varepsilon_1(\lambda_2-qa_1\varepsilon_1)}{1-qa_2\varepsilon_2}\times
\frac{1-(qa_2\varepsilon_2)^{n-1}}{1-qa_2\varepsilon_2}\nonumber\\
& & -\frac{2}{(\lambda_2-\lambda_1)(1-\lambda_1)}\left(
\frac{qa_1\varepsilon_1(\lambda_2-qa_1\varepsilon_1)}{1-qa_2\varepsilon_2}-qa_2\varepsilon_2
\right)\frac{\lambda_1(1-\lambda_1^{n-1})}{1-\lambda_1}\nonumber\\
& & + \frac{2}{(\lambda_2-\lambda_1)(1-\lambda_1)}\times
\frac{qa_1\varepsilon_1(\lambda_2-qa_1\varepsilon_1)}{1-qa_2\varepsilon_2}\times
\frac{\lambda_1(\lambda_1^{n-1}-(qa_2\varepsilon_2)^{n-1})}{\lambda_1-qa_2\varepsilon_2}
\nonumber\\
& & + \frac{2(n-1)}{(\lambda_2-\lambda_1)(1-\lambda_2)}\left(
\frac{qa_1\varepsilon_1(qa_1\varepsilon_1-\lambda_1)}{1-qa_2\varepsilon_2}+qa_2\varepsilon_2
\right)\nonumber\\
& & - \frac{2}{(\lambda_2-\lambda_1)(1-\lambda_2)}\times
\frac{qa_1\varepsilon_1(qa_1\varepsilon_1-\lambda_1)}{1-qa_2\varepsilon_2}\times
\frac{1-(qa_2\varepsilon_2)^{n-1}}{1-qa_2\varepsilon_2}\nonumber\\
& & -\frac{2}{(\lambda_2-\lambda_1)(1-\lambda_2)}\left(
\frac{qa_1\varepsilon_1(qa_1\varepsilon_1-\lambda_1)}{1-qa_2\varepsilon_2}+qa_2\varepsilon_2
\right)\frac{\lambda_2(1-\lambda_2^{n-1})}{1-\lambda_2}\nonumber\\
& & + \frac{2}{(\lambda_2-\lambda_1)(1-\lambda_2)}\times
\frac{qa_1\varepsilon_1(qa_1\varepsilon_1-\lambda_1)}{1-qa_2\varepsilon_2}\times
\frac{\lambda_2(\lambda_2^{n-1}-(qa_2\varepsilon_2)^{n-1})}{\lambda_2-qa_2\varepsilon_2}
\end{eqnarray*}
\end{small}
For our purpose, a complete analytic discussion of this expression
would take us too far. Instead, we will illustrate the behavior of
the variance with four typical choices for the set of the parameters
$\varepsilon_1,\varepsilon_2,a_1,a_2$ in Figures~\ref{fig:PM2a5m1p5m5++},
\ref{fig:PM2a5m1p5m5+-}, \ref{fig:PM2a5m1p5m5-+}, and
\ref{fig:PM2a5m1p5m5--}. Note that we obtain crossovers, respectively,
similar to those obtained for the persistent and antipersistent walks
of the Taylor's model when $\varepsilon_2=1$ and $\varepsilon_1$ is either equal
to $1$ or $-1$. On the contrary when $\varepsilon_2=-1$, There is no crossover,
the variance behaves as $n$ for both signs of $\varepsilon_1$.
\vspace{0.3cm}
\begin{figure}[h]
\noindent
\begin{minipage}{0.46\linewidth}
\centering\epsfig{figure=LM2a5m1p5m5++.eps, width=\linewidth}
\caption{\label{fig:PM2a5m1p5m5++}\textit{Log-log plot of $<R_n^2>$ vs.n
for $M=2$, $\varepsilon_1=1$, $\varepsilon_2=1$, $a_1=a_2=0.5$, $p=5\,10^{-5}$,
$n_{\max}=1.6\,10^8$.}}
\end{minipage}\hfill
\begin{minipage}{0.46\linewidth}
\centering\epsfig{figure=LM2a5m1p5m5+-.eps, width=\linewidth}
\caption{\label{fig:PM2a5m1p5m5+-}\textit{Log-log plot of $<R_n^2>$ vs.n
for $M=2$, $\varepsilon_1=1$, $\varepsilon_2=-1$, $a_1=a_2=0.5$, $p=5\,10^{-5}$,
$n_{\max}=1.6\,10^8$.}}
\end{minipage}
\end{figure}
\vspace{0.3cm}
\begin{figure}[h]
\noindent
\begin{minipage}{0.46\linewidth}
\centering\epsfig{figure=LM2a5m1p5m5-+.eps, width=\linewidth}
\caption{\label{fig:PM2a5m1p5m5-+}\textit{Log-log plot of $<R_n^2>$ vs.n
for $M=2$, $\varepsilon_1=-1$, $\varepsilon_2=1$, $a_1=a_2=0.5$, $p=5\,10^{-5}$,
$n_{\max}=1.6\,10^8$.}}
\end{minipage}\hfill
\begin{minipage}{0.46\linewidth}
\centering\epsfig{figure=LM2a5m1p5m5--.eps, width=\linewidth}
\caption{\label{fig:PM2a5m1p5m5--}\textit{Log-log plot of $<R_n^2>$ vs.n
for $M=2$, $\varepsilon_1=-1$, $\varepsilon_2=-1$, $a_1=a_2=0.5$, $p=5\,10^{-5}$,
$n_{\max}=1.6\,10^8$.}}
\end{minipage}
\end{figure}
\subsection{Linear memory for $M>2$.}
The matrix method used for the case $M=2$ is formally easy to generalize. It
yields, for the generating function, a recursion relation similar to
\ref{2fiter}, where $\mathbf{f}_n$ is a $2^M$-component vector,
$\mathbf{M}$ a $2^M\times 2^M$ matrix and $\mathcal{P}_4$ is replaced by a
$2^M\times 2^M$ permutation matrix $\mathcal{P}_{2^M}$. With increasing values
of $M$ it becomes rapidly untractable. Unfortunately, due to the large number
of parameters when $M>2$, even the variance $<R_n^2>$ is unmanageable.
This is not the case for the single product memory, as will be seen
in the next section. However, there is, at least, one case for which we can
predict that the variance exhibits the crossover observed in the persistent
walk of the Taylor's model. This occurs when all the $\varepsilon_j$
($j=1,2,\ldots,M$) are equal to 1, since in this case, it is easy to verify
that for $p=0$, the variance behaves as $n^2$.
\section{Single product memory function.}
Here again we will first consider the case $M=2$ as an illustration
of the complexity which manifests as soon as $M>1$. Since we will
only consider the case of a single product, the notation
$\varepsilon_{j_1j_2\cdots j_r}$ will be replaced by $\varepsilon$.
\subsection{Probability distribution for $M=2$.}
As in the linear memory function case,
$p_n^{\sigma_1\sigma_2}(k)$ denotes the probability
$P((R_n=k)\cap((S_{n-1}=\sigma_1)\cap(S_n=\sigma_2)))$,
and as above, if $S_{n-1}=\sigma_1$ and $S_n=\sigma_2$, then
$R_{n-1}=k-\sigma_2$, which implies that $p_n^{\sigma_1\sigma_2}(k)$
is a linear combination of the two probabilities
$p_{n-1}^{\sigma\sigma_1}(k-\sigma_2)$, where $\sigma = \pm 1$.
Since, with a probability $q$, $S_n=\varepsilon S_{n-1}S_{n-2}$, we have
$\sigma_2=\varepsilon\sigma\sigma_1$ or $\sigma=\varepsilon\sigma_1\sigma_2$.
Therefore, using notations similar to those of the linear case, we have
$$
\left(\begin{array}{c}
f_n^{++}(x)\\[0.2cm]
f_n^{+-}(x)
\end{array}\right)
=
\mathbf{M^+}
\left(\begin{array}{c}
f_{n-1}^{++}(x)\\[0.2cm]
f_{n-1}^{-+}(x)
\end{array}\right),
$$
where
\begin{equation}
\mathbf{M^+}(x) = \left(\begin{array}{cc}
\frac{1}{2}
(1+q\varepsilon)\; x
& \frac{1}{2}
(1-q\varepsilon)\;x\\[0.2cm]
\frac{1}{2}
(1-q\varepsilon)\; \frac{1}{x}
&\frac{1}{2}
(1+q\varepsilon)\;\frac{1}{x}
\end{array}\right),
\label{PMplus}
\end{equation}
and
$$
\left(\begin{array}{c}
f_n^{-+}(x)\\[0.2cm]
f_n^{--}(x)
\end{array}\right)
=
\mathbf{M^-}
\left(\begin{array}{c}
f_{n-1}^{+-}(x)\\[0.2cm]
f_{n-1}^{-+}(x)
\end{array}\right),
$$
where
\begin{equation}
\mathbf{M^-}(x) = \left(\begin{array}{cc}
\frac{1}{2}
(1-q\varepsilon)\; x
& \frac{1}{2}
(1+q\varepsilon)\;x\\[0.2cm]
\frac{1}{2}
(1+q\varepsilon)\; \frac{1}{x}
&\frac{1}{2}
(1-q\varepsilon)\;\frac{1}{x}
\end{array}\right).
\label{PMminus}
\end{equation}
The above equations can be grouped together in a unique
equation that reads
$$
\left(\begin{array}{c}
f_n^{++}(x)\\[0.18cm]
f_n^{+-}(x)\\[0.18cm]
f_n^{-+}(x)\\[0.18cm]
f_n^{--}(x)
\end{array}\right)
=
\left(\begin{array}{cc}
\mbox{\huge{$\mathbf{M^+}(x)$}} & \mbox{\huge{$\mathbf{0}$}}\\[0.8cm]
\mbox{\huge{$\mathbf{0}$}} & \mbox{\huge{$\mathbf{M^-}(x)$}}
\end{array}\right)
\left(\begin{array}{c}
f_{n-1}^{++}(x)\\[0.18cm]
f_{n-1}^{-+}(x)\\[0.18cm]
f_{n-1}^{+-}(x)\\[0.18cm]
f_{n-1}^{--}(x)
\end{array}\right).
$$
Since the column vector on the right hand side of the above equation is not
correctly ordered, we have to introduce a $4\times 4$ permutation matrix, and
write this equation under the final form
\begin{equation}
\mathbf{f}_n(x) = \mathbf{M}(x)\;\mathcal{P}_4\; \mathbf{f}_{n-1}(x)
\label{P2fiter}
\end{equation}
where
$$
\mathbf{f}_n(x) =
\left(\begin{array}{c}
f_n^{++}(x)\\[0.2cm]
f_n^{+-}(x)\\[0.2cm]
f_n^{-+}(x)\\[0.2cm]
f_n^{--}(x)\\
\end{array}\right),
$$
$$
\mathbf{M}(x) =
\left(\begin{array}{cc}
\mathbf{M^+}(x) & \mathbf{0}\\[0.2cm]
\mathbf{0} & \mathbf {M^-}(x)
\end{array}\right)
$$
and
$$
\mathcal{P}_4=
\left(\begin{array}{cccc}
1 & 0 & 0 & 0\\
0 & 0 & 1 & 0\\
0 & 1 & 0 & 0\\
0 & 0 & 0 & 1
\end{array}\right).
$$
Iterating Relation~\ref{P2fiter}, we obtain
\begin{equation}
\mathbf{f}_n(x)
= \left(\mathbf{M}(x)\;\mathcal{P}_4\right)^{n-2} \; \mathbf{f}_2(x),
\label{recur-P2fn}
\end{equation}
where
$$
\mathbf{f}_2(x) =
\left(\begin{array}{c}
f_{2}^{++}(x)\\[0.2cm]
f_{2}^{+-}(x)\\[0.2cm]
f_{2}^{-+}(x)\\[0.2cm]
f_{2}^{--}(x)\\
\end{array}\right)=
\left(\begin{array}{c}
\frac{x^2}{4}\\[0.2cm]
\frac{1}{4}\\[0.2cm]
\frac{1}{4}\\[0.2cm]
\frac{1}{4x^2}
\end{array}\right).
$$
\vspace{0.3cm}
\begin{figure}[b]
\noindent
\begin{minipage}{0.46\linewidth}
\centering\epsfig{figure=PM2n50p5m3.eps, width=\linewidth}
\caption{\label{fig:PM2n50p5m3}\textit{$p_n(k)$ for
$M=2$, $\varepsilon=1$, $p=0.005$, and $n=50$.}}
\end{minipage}\hfill
\begin{minipage}{0.46\linewidth}
\centering\epsfig{figure=PM2n50p5m1.eps, width=\linewidth}
\caption{\label{fig:PM2n50p5m1}\textit{$p_n(k)$ for
$M=2$, $\varepsilon=1$, $p=0.5$, and $n=50$.}}
\end{minipage}
\end{figure}
\vspace{0.3cm}
\begin{figure}[t]
\noindent
\begin{minipage}{0.46\linewidth}
\centering\epsfig{figure=PM2n50p5m3-.eps, width=\linewidth}
\caption{\label{fig:PM2n50p5m3-}\textit{$p_n(k)$ for
$M=2$, $\varepsilon=-1$, $p=0.005$, and $n=50$.}}
\end{minipage}\hfill
\begin{minipage}{0.46\linewidth}
\centering\epsfig{figure=PM2n50p5m1-.eps, width=\linewidth}
\caption{\label{fig:PM2n50p5m1-}\textit{$p_n(k)$ for
$M=2$, $\varepsilon=-1$, $p=0.5$, and $n=50$.}}
\end{minipage}
\end{figure}
As in the linear case, further progress toward the expression of the generating
function $f_n(x)$ is extremely complicated, and would
reveal untractable for $M>2$.
However the formalism is well suited to numerically evaluate
the probability distribution of $R_n$ recursively using~\ref{recur-P2fn}.
Figures~\ref{fig:PM2n50p5m3}, \ref{fig:PM2n50p5m1}, \ref{fig:PM2n50p5m3-},
\ref{fig:PM2n50p5m1-} show the probability distribution $p_n(k)$
for two values of $p$ to illustrate cases $np\ll 1$ and $np\gg 1$ when
$\varepsilon=\pm1$.
At the light of the comment at the end of subsection 3.0.1,
Figures~\ref{fig:PM2n50p5m3}, \ref{fig:PM2n50p5m1}, \ref{fig:PM2n50p5m3-},
\ref{fig:PM2n50p5m1-} reveal the existence of crossovers. This property
will be confirmed directly in the next subsection.
\subsection{Variance for $M=2$.}
In order to evaluate recursively the variance, we need to determine
the correlations. We have
\begin{eqnarray*}
<S_\ell S_{\ell+1}> & = & \varepsilon q<S_\ell (S_{\ell}S_{\ell-1})>\\
& = & \varepsilon q <S_{\ell-1}>=0;
\end{eqnarray*}
\begin{eqnarray*}
<S_\ell S_{\ell+2}> & = & \varepsilon q<S_\ell (S_{\ell+1}S_\ell)>\\
& = & \varepsilon q <S_{\ell+1}>=0;
\end{eqnarray*}
\begin{eqnarray*}
<S_\ell S_{\ell+3}> & = & \varepsilon q<S_\ell (S_{\ell+2}S_{\ell+1})>\\
& = & \varepsilon^2 q^2 <S_\ell((S_{\ell+1}S_\ell)S_{\ell+1}>\\
& = & q^2;
\end{eqnarray*}
and, for $m>\ell+3$,
\begin{eqnarray*}
<S_\ell S_m> & = & \varepsilon q<S_\ell (S_{m-1}S_{m-2})>\\
& = & \varepsilon^2 q^2 <S_\ell((S_{m-2}S_{m-3})S_{m-2}>\\
& = & q^2 <S_\ell S_{m-3}>.
\end{eqnarray*}
These results provide all the correlations:
\begin{equation}
<S_\ell S_m> =
\begin{cases}
0 & \text{if $m-\ell$ is not a multiple of 3},\\[0.1cm]
q^{2(m-\ell)/3} & \text{if $m-\ell$ is a multiple of 3}.
\end{cases}
\label{correlPM2}
\end{equation}
Note that the correlations do not depend upon $\varepsilon$.
On the other hand, the variance satisfies the recursion relation
$$
<R_n^2> = <R_{n-1}^2> + 1 + 2<S_1S_n + S_2S_n + \cdots + S_{n-1}S_n>.
$$
Iterating this relation and taking into account \ref{correlPM2}, we
finally obtain
\begin{eqnarray}
<R_n^2> & = & n + (n-3)q^2+(n-6)q^4+\cdots+
(n-3\lfloor n/3\rfloor) q^{2\lfloor n/3\rfloor}\nonumber\\
& = & n + n q^2\frac{1-q^{2\lfloor n/3\rfloor}}{1-q^2}\nonumber\\
& &\quad - 3q^2\frac{\lfloor n/3\rfloor q^{2\lfloor n/3\rfloor+2}-
\left(\lfloor n/3\rfloor+1\right)q^{2\lfloor n/3\rfloor} +1}
{(1-q^2)^2},\label{varPM2}
\end{eqnarray}
where $\lfloor x\rfloor$ denotes the largest integer less or equal to $x$.
On this expression of $<R_n^2>$ we readily verify that for $q=0$ the
variance is equal to $n$, while for $q=1$ it takes the value
$$
n + (2n-3)\lfloor n/3\rfloor - 3 (\lfloor n/3\rfloor)^2 = \lceil n^2/3\rceil,
$$
where $\lceil x\rceil$ denotes the smallest integer greater or equal to $x$.
Since the functions floor and ceiling are not differentiable, we define the
crossover exponent $E(n,p)$ by
$$
E(n,p)= \frac{n}{2}\frac{<R_{n+1}^2>(p)-<R_{n-1}^2>(p)}{<R_n^2>(p)}.
$$
Figure~\ref{fig:PM2H5m5} represents the variation of $E(n,p)$ as a function of
of $\log n$ for $p=5\,10^{-5}$.
\vspace{0.3cm}
\begin{figure}[h]
\centerline{\includegraphics*[scale=0.8]{PM2H5m5.eps}}
\caption{\label{fig:PM2H5m5}\textit{Exponent $E(n,p)$ as a function of
$\log n$, for $M=2$, $p=5\,10^{-5}$ and $n_{\max}=1.6\,10^7$.}}
\end{figure}
\subsection{Probability distribution for $M>2$.}
The matrix method used for the case $M=2$ is formally easy to generalize. It
yields, for the generating function, a recursion relation similar to
\ref{P2fiter}, where $\mathbf{f}_n$ is a $2^M$-component vector,
$\mathbf{M}$ a $2^M\times 2^M$ matrix and $\mathcal{P}_4$ is replaced by a
$2^M\times 2^M$ permutation matrix $\mathcal{P}_{2^M}$. With increasing values
of $M$ it becomes rapidly untractable. However, the variance $<R_n^2>$ can be
exactly evaluated as we will show in next subsection.
\subsection{Variance for $M>2$.}
In order to evaluate recursively the variance, we need to determine
the correlations. By definition, $<S_\ell S_m>=0$ if $\ell<m\le M$. This is
also true if $m=M+1$:
\begin{eqnarray*}
<S_\ell S_{M+1}> & = & q\varepsilon<S_\ell(S_1S_2\cdots S_\ell\cdots S_M)>\\
& = &q\varepsilon <S_1><S_2>\cdots<S_M>=0,
\end{eqnarray*}
due the fact that the random variables $S_1, S_2, \ldots, S_M$
are independent, and $<S_\ell^2>=1$.
Considering now the case: $<S_\ell S_m>$ (any $\ell$ and $m>\ell$), we first
note that
\begin{eqnarray*}
<S_\ell S_{\ell+M+1}> & = & q\varepsilon<S_\ell(S_{\ell+1}S_{\ell+2}\cdots S_{\ell+M})>\\
& = & (q\varepsilon)^2<(S_\ell S_{\ell+1}\cdots S_{\ell+M-1})^2>\\
& = & q^2,
\end{eqnarray*}
whereas, in the general case,
\begin{eqnarray}
<S_\ell S_{\ell+r}> & = & q\varepsilon
<S_\ell(S_{\ell+r-M}S_{\ell+r-M+1}\cdots S_{\ell+r-1})>\nonumber\\
& = & q^2<S_\ell S_{\ell+r-(M+1)}>,
\label{correliterPM}
\end{eqnarray}
where it is assumed that $\ell+r-(M+1)$ is positive but may be smaller
than $\ell$. Iterated application of~\ref{correliterPM} in the presence
of the initial conditions shows that, for
$m-\ell = \big\lfloor{\frac{m-\ell}{M+1}}\big\rfloor + b$ ($0\leq b<M+1$),
\begin{equation}
<S_\ell S_m> =
\begin{cases}
0 & \text{if $b>0$}\\
q^{2\big\lfloor\frac{m-\ell}{M+1}\big\rfloor} & \text{if $b=0$}\\
\end{cases}
\label{correlPMgeneral}
\end{equation}
Substitution of~\ref{correlPMgeneral} in the recursion relation satisfied by
the variance:
$$
<R_n^2> = <R_{n-1}^2> + 1 + 2<S_1S_n + S_2S_n + \cdots + S_{n-1}S_n>.
$$
yields
\begin{eqnarray}
<R_n^2> & = & n + (n-(M+1))q^2 + (n -2(M+1))q^4+\cdots\nonumber\\
& & +\left(n-\Big\lfloor\frac{n}{M+1}\Big\rfloor(M+1)\right)
q^{2\lfloor\frac{n}{M+1}\rfloor}\nonumber\\
& = & n + n q^2\frac{1-q^{2\lfloor\frac{n}{M+1}\rfloor}}{1-q^2}
-(M+1)q^2\nonumber\\
& & \times\frac{1 - \left(\big\lfloor\frac{n}{M+1}\big\rfloor+1\right)
q^{2\lfloor\frac{n}{M+1}\rfloor}+\big\lfloor\frac{n}{M+1}\big\rfloor
q^{2\lfloor\frac{n}{M+1}\rfloor+2}}{(1-q^2)^2},\label{varPMofq}
\end{eqnarray}
which is the generalization for any $M$ of the result~\ref{varPM2}
obtained for $M=2$. Note that $M=1$ is not a particular case of~\ref{varPMofq}.
When $q=0$, $<R_n^2>=n$, whereas, when $p$ goes to zero ($q\to 1$), $<R_n^2>$
tends to
$$
n + (2n-(M+1))\bigg\lfloor\frac{n}{M+1}\bigg\rfloor,
$$
which is of the order of $\frac{n^2}{M+1}$ for large $n$.
These two limiting behaviors are the signal of a crossover, which is evidenced
in Figure~\ref{fig:multiPMVp5m5}. This figure shows that the variance is a
decreasing function of $M$ at a given $n$.
\vspace{0.3cm}
\begin{figure}[h]
\centerline{\includegraphics*[scale=0.8]{multiPMVp5m5.eps}}
\caption{\label{fig:multiPMVp5m5}\textit{Log-log plot of the variance
as a function of $n$,
for $M=2,8,32,128$ (to be read downwards), $p=5\,10^{-5}$ and
$n_{\max}=1.6\,10^8$.}}
\end{figure}
The general expression of the variance is an even function of $q$, and,
therefore, does not depend upon $\varepsilon$. Such is not the case for the
probability distribution $p_n(k)$ of $R_n$ as in Figures~\ref{fig:PM5n13p0+} and
\ref{fig:PM5n13p0-}.
\vspace{0.3cm}
\begin{figure}[h]
\noindent
\begin{minipage}{0.46\linewidth}
\centering\epsfig{figure=PM5n13p0+.eps, width=\linewidth}
\caption{\label{fig:PM5n13p0+}\textit{$p_n(k)$ for
$M=5$, $\varepsilon=1$, $p=0$, and $n=13$. $<R_{13}^2>=29$.}}
\end{minipage}\hfill
\begin{minipage}{0.46\linewidth}
\centering\epsfig{figure=PM5n13p0-.eps, width=\linewidth}
\caption{\label{fig:PM5n13p0-}\textit{$p_n(k)$ for
$M=5$, $\varepsilon=-1$, $p=0$, and $n=13$. $<R_{13}^2>=29$.}}
\end{minipage}
\end{figure}
Here the probability distribution are symmetric. This is not always the case as
can be observed in Figures~\ref{fig:PM2n50p5m3}, \ref{fig:PM2n50p5m1},
\ref{fig:PM2n50p5m3-}, and \ref{fig:PM2n50p5m1-}.
\section{Conclusion and perspectives}
We have studied a family of correlated random walks with a finite memory
range. These walks are a generalization of the Taylor's walk.
We have established a matrix formalism for the generating functions of
the probability distributions of these walks. We have also derived in many
cases an analytic expression for the variance. As a function of the number
of steps $n$, for each of these walks, the variance becomes ultimately
linear in $n$ in the limit $n\to\infty$. However, at comparatively small $n$,
a different behavior of the variance can be observed leading to the existence
of a crossover phenomenon. This feature shows that, when analyzing experimental
results, one finds a Hurst exponent different from $\frac{1}{2}$, this does not
necessarily imply an infinite-range memory. If, experimentally feasible, one
should increase significantly the number of steps to cross-check the stability
of the exponent before concluding that the system under consideration exhibits
a non-Gaussian behavior.
The ultimate Gaussian behavior of the variance
at very large $n$ is a direct consequence of the exponential decay of the
step-step correlations as a function of the absolute value of the difference
between step indices. Therefore, in order to find a different behavior, one
has to consider step-step correlations decreasing more slowly than
exponentially. This is what may happen in the case of a power law decay
as can be shown as follows.
For a symmetric random walk, the variance is given by
\begin{eqnarray*}
<R_n^2> & = & n + 2 \sum_{\ell=1}^{n-1}\sum_{m=\ell+1}^n <S_\ell S_m>\\
& = & n + 2 \sum_{\ell=1}^{n-1}\sum_{r=1}^n <S_\ell S_{\ell+r}>\\
& = & n + 2 \sum_{\ell=1}^{n-1}\sum_{r=1}^n c_{r,\ell},
\end{eqnarray*}
where as above $c_{r,\ell}=<S_\ell S_{r+\ell}>$. If we now assume that,
for some reason,
$$
c_{r,\ell} = \frac{1}{r^\alpha}
$$
independently of $\ell$, then
$$
<R_n^2> = n + 2\sum_{r=1}^{n-1}\frac{n-r}{r^\alpha}.
$$
The possible non-Gaussian behavior of $<R_n^2>$ can only result from the
asymptotic behavior at large $n$ of the summation in the right hand side
of the above relation. According to the value of $\alpha$, the
asymptotic behaviors of
$$
S(n,\alpha) = \sum_{r=1}^{n-1}\frac{n-r}{r^\alpha}
$$
depends on $\alpha$ as follows:
\begin{itemize}
\item{if $\alpha>1$, $S(n,\alpha)\sim n\zeta(\alpha)$, where $\zeta$ is the Riemann
function.}
\item{if $\alpha=1$, $S(n,1)\sim n\log n$.}
\item{if $\alpha<1$, $S(n,\alpha)\sim\frac{1}{(1-\alpha)(2-\alpha)}n^{2-\alpha}$.}
\end{itemize}
Note that the power law decay for the step-step correlations is only a
sufficient condition to obtain a non-trivial Hurst exponent equal to $1-a/2$.
\section*{References}
{\parindent = 0pt
[1] For a review of early work on random walk, see W. Feller,
An Introduction to Probability Theory and Its Applications,
John Wiley and Sons, New York, 1960, Volume I, Chapter XIV.
[2] J. J. Collins and C. J. De Luca, Random Walking during Quiet
Standing, Phys. Rev. Lett., {\bf 73}, 764--767, 1994.
[3] L. Nunes Amaral, A. Goldberger, P. Ivanov and E. Stanley, Scale-Indepen\-dent
Measures and Pathologic Cardiac Dynamics, Phys. Rev. Lett., {\bf 81},
2388--2391, 1998.
[4] W. Li, The Study of Correlated Structures of DNA Sequences: A Critical
Review, lanl archives, adap-org/9704003
[5] C.-K. Peng, S. Buldyrev, A. Goldberger, S. Havlin, F. Sciortino,
M. Simons and E. Stanley, Long-range correlations in nucleotide sequences,
Nature, {\bf 356}, 168--170, 1992.
[6] G. Abramson, P. Alemany and H. Cerdeira, Noisy L\'evy walk analog
of two-dimensional DNA walks for chromosomes of S. cerevisiae, Phys. Rev. E
{\bf 58,} 914--918, 1998.
[7] L. Liebovitch and W. Yang, Transition from persistent to antipersistent
correlation in biological systems, Phys. Rev. E, {\bf 56}, 4557--4566, 1997.
[8] P. Ivanov, L. Nunes Amaral, A. Goldberger and E. Stanley, Stochastic
feedback and the regulation of biological rhythms, Europhys. Lett., {\bf 43},
363--368, 1998.
[9] R. Mantegna and E. Stanley, Scaling behaviour in the dynamics of an
economic index, Nature, {\bf 376}, 46--49, 1995.
[10] R. Mantegna and E. Stanley, Turbulence and financial markets, Nature,
{\bf 383}, 587--588, 1996.
[11] E. Scalas, Scaling in the market of futures, Physica A, {\bf 253},
394--402, 1998.
[12] G. I. Taylor, Diffusion by continuous movements, Proceedings
of the London Mathematical Society, {\bf 20}, 196--212, 1921/22.
[13] S. Goldstein, On diffusion by discontinuous movements, and the
telegraph equation, The Quarterly Journal of Mechanics and Applied
Mathematics {\bf 4}, 129--156, 1951.
[14] J. Gillis, Correlated random walk, Proc. Camb. Phys. Soc.
{\bf 51}, 639--651, 1955.
[15] E. Renshaw and R. Henderson, The correlated random walk, Journal
of Applied Probability, {\bf 18}, 403--414, 1981.
[16] A. Chen and E. Renshaw, The Gillis-Domb-Fisher correlated random
walk, Journal of Applied Probability, {\bf 29}, 792--813, 1992.
[17] G. H. Weiss, Aspects and Applications of the random walk,
North-Holland, Amsterdam, 1994.
}
\end{document}
\begin{eqnarray*}
p^+(t,x) & = & \left(1-\frac{p}{2}\right) \; p^+(t-\tau, x-\xi) +
\frac{p}{2} \; p_-(t-\tau, x-\xi)\\
p^-(t,x) & = & \frac{p}{2} \; p^+(t-\tau, x+\xi) +
\left(1-\frac{p}{2}\right) \; p^-(t-\tau, x+\xi)\\
\end{eqnarray*}
\begin{eqnarray*}
p^+(t,x) & = & \left(1-\frac{p}{2}\right) \;\left(p^+(t,x) -
\frac{\partial p^+}{\partial t}\;\tau -
\frac{\partial p^+}{\partial x}\;\xi +
\frac{1}{2}\;\frac{\partial^2 p^+}{\partial x^2}\;\xi^2\right)\\
& & \qquad +
\frac{p}{2} \; \left(p^-(t,x) -
\frac{\partial p^-}{\partial t}\;\tau -
\frac{\partial p^-}{\partial x}\;\xi +
\frac{1}{2}\;\frac{\partial^2 p^+}{\partial x^2}\;\xi^2\right)\\
p^-(t,x) & = & \frac{p}{2} \; \left(p^+(t,x) -
\frac{\partial p^+}{\partial t}\;\tau +
\frac{\partial p^+}{\partial x}\;\xi +
\frac{1}{2}\;\frac{\partial^2 p^+}{\partial x^2}\;\xi^2\right)\\
& & \qquad +
\left(1-\frac{p}{2}\right) \; \left(p_-(t,x) -
\frac{\partial p^-}{\partial t}\;\tau +
\frac{\partial p^-}{\partial x}\;\xi +
\frac{1}{2}\;\frac{\partial^2 p^+}{\partial x^2}\;\xi^2\right)\\
\end{eqnarray*}
$$
\begin{array}{ccccccc}
<R_n^2> & = & <R_{n-1}^2> & + & 1 & + & 2<S_1S_n + S_2S_n
+ \cdots + S_{n-1}S_n>\\
<R_{n-1}^2> & = & <R_{n-2}^2> & + & 1 & + & 2<S_1S_{n-1}
+ \cdots + S_{n-2}S_{n-1}>\\
& \cdots & \\
<R_2^2> & = & <R_1^2> & + & 1 & + & 2 <S_1S_2>\\
<R_1^2> & = & & & 1 &
\end{array}
$$
|
\section{Introduction}
Due to holomorphicity properties and non\discretionary{-}{}{-}re\-nor\-ma\-li\-za\-tion\ theorems valid in
$N=1$ supersymmetric theories it has become possible to argue that some
supersymmetric gauge theories with special matter content confine at low
energies. The first example is due to Seiberg \cite{Seib_D49} who found
that supersymmetric quantum chromodynamics (SQCD) with gauge group $SU(N_c)$
and $N_f$ quark flavors shows confinement when $N_f=N_c$ or $N_f=N_c+1$.
This has been generalized to more complicated models. All $N=1$
supersymmetric gauge theories with vanishing tree-level superpotential
which confine at low energies could be classified \cite{confine,qmm,
affineqm} because they are constrained by an index argument.
When a tree-level superpotential is present the index argument is no longer
valid. Because of the lower symmetry the non-perturbative superpotential is
less constrained and one expects more confining models to exist. Indeed,
Cs\'aki and Murayama \cite{newconf} showed that many of the Kutasov-like
\cite{KuSch} models exhibit confinement for special values of the number
of quark flavors $N_f$.\footnote{Some further confining models with non\discretionary{-}{}{-}van\-ish\-ing\
tree-level superpotential are discussed in \cite{TaYo}.}
These models contain fields in tensor representations of the gauge group
and an appropriate superpotential for these tensor fields. For all of
the models considered in \cite{newconf} a dual description in terms of
magnetic variables is known \cite{KuSch,Intri,LS,ILSdual} and the authors of
\cite{newconf} used the fact that the electric gauge theory confines when
its magnetic dual is completely higgsed.
Seiberg already used this idea as an additional consistency check in his
original paper establishing e\-lec\-tric\discretionary{-}{}{-}mag\-ne\-tic\ duality for non\discretionary{-}{}{-}Abel\-ian\
$N=1$ supersymmetric gauge theories \cite{Seib}. He showed how the confining
superpotential of SQCD with $N_f=N_c+1$ could be obtained by a perturbative
calculation in the completely broken magnetic gauge theory. Under duality the
fields of the magnetic theory (which are gauge singlets as the gauge
symmetry is completely broken) are mapped to the mesons and baryons of
the electric theory and the confining superpotential is easily shown
to be the image of the magnetic superpotential under this mapping \cite{Seib}.
This a realization in $N=1$ supersymmetric gauge theories of an old idea of
't Hooft and Mandelstam \cite{dualmeiss} that confinement is driven by
condensation of magnetic monopoles.
Now, many gauge theory models have been found that possess a dual description
in terms of magnetic variables in the infrared. This allows us to predict
many new examples of confining gauge theories. The idea described in the
previous paragraph was first used by the authors of \cite{KSS} to determine
the confining spectrum of the model proposed by Kutasov \cite{KuSch} and has
been applied by Cs\'aki and Murayama \cite{newconf} to six further
models that confine in the presence of an appropriate superpotential.
In this talk I review the original example of Seiberg \cite{Seib} and
explain how e\-lec\-tric\discretionary{-}{}{-}mag\-ne\-tic\ duality is used to obtain the low-energy
spectrum and the form of the non\discretionary{-}{}{-}per\-tur\-ba\-tive\ superpotential of confining gauge
theories \cite{mklein}. One finds that all of the gauge theory models based
on simple gauge groups considered in \cite{ILSdual,patterns} confine when the
gauge groups of their magnetic duals are completely broken by the Higgs
effect. For nine of these theories the confining phase has not been discussed
before.
\section{Phase structure of SQCD}
Let us briefly review the well-known phase structure of SQCD \cite{ISlectures}.
By this we mean an $N=1$ supersymmetric $SU(N_c)$ gauge theory with $N_f$
quark flavors, i.e.\ $N_f$ chiral matter supermultiplets $Q$ transforming in
the fundamental representation of the gauge group and the same amount of matter
multiplets $\bar Q$ transforming in the antifundamental\ representation.
Consider first the case of vanishing tree-level superpotential. Depending on
the relative values of $N_f$ and $N_c$ the the low-energy theory resides in
different phases.
$N_f=0$: This is pure super Yang-Mills theory. It is believed to show
confinement. According to an index argument by Witten \cite{W-index} there
are $N_c$ distinct supersymmetric vacua.
$0<N_f<N_c$: There is a non\discretionary{-}{}{-}per\-tur\-ba\-tive\ superpotential generated by gluino
condensation (for $N_f<N_c-1$) or by instantons (for $N_f=N_c-1$),
as was shown by Affleck, Dine and Seiberg \cite{ADS}:
\begin{equation} \label{WADS}
W_{\rm np}=(N_c-N_f)\left(\Lambda^{3N_c-N_f}\over\det M\right)^{1\over N_c-N_f}\,,
\end{equation}
where $\Lambda$ is the dynamically generated scale of the theory and the
meson matrix $M$ is defined by $M^{ij}=Q^{\alpha i}\bar Q^j_\alpha$,
$i,j=1,\ldots,N_f$, $\alpha=1,\ldots,N_c$. The minimum of the potential lies
at infinite field expectation values and therefore the theory has no stable
vacuum for this range of parameters.
$N_f=N_c$: In this case the superpotential vanishes even at the non\discretionary{-}{}{-}per\-tur\-ba\-tive\ level
\cite{Seib_D49}. As a consequence the flat directions that pa\-ram\-e\-trize
the moduli space of vacua are not lifted in the quantum theory. The low-energy
spectrum is given by the mesons $M^{ij}$ defined above and the baryons
$B=\det Q$, $\bar B=\det\bar Q$ (the quarks $Q$, $\bar Q$ are viewed as
$(N_f\times N_c)$-matrices). The physical degrees of freedom at low energies
being gauge invariant means that the theory confines. The classical
constraint $\det M=B\bar B$ is modified in the quantum theory \cite{Seib_D49}
to
\begin{equation} \label{qconstr}
\det M - B\bar B = \Lambda^{2N_c}\,.
\end{equation}
The expectation values of the mesons and baryons that satisfy this constraint
span the quantum moduli space. The observation that the expectation values
of $\det M$ and $B\bar B$ cannot vanish simultaneously tells us that the
chiral symmetry is spontaneously broken.
$N_f=N_c+1$: There is again a quantum moduli space, but now the classical
constraints are not modified by quantum effects. In the low-energy theory
they can be derived from the non\discretionary{-}{}{-}per\-tur\-ba\-tive\ superpotential \cite{Seib_D49}
\begin{equation} \label{Wconf}
W_{\rm np}={\bar BMB-\det M\over\Lambda^{2N_f-3}}\,,
\end{equation}
where the baryons $B^i$ are defined as the determinant of the quark matrix $Q$
with the $i$-th line omitted. This describes confinement without \linebreak
breaking of the chiral symmetry.
$N_c+2\le N_f\le {3\over2}N_c$: The low-energy energy theory is rather
complicated and more appropriately described in terms of dual magnetic
variables. The dual magnetic theory is infrared free for this range of
parameters.
${3\over2}N_c<N_f<3N_c$: At low energies the theory is driven to an infrared
fixed point of the renormalization group \cite{Seib} and resides in a non\discretionary{-}{}{-}Abel\-ian\
Coulomb phase. Seiberg found a dual description of this model \cite{Seib} by
an $SU(N_f-N_c)$ gauge theory with $N_f$ (magnetic) quark flavors $q$, $\bar q$
and $N_f^2$ additional singlets $M_{\rm mag}^{ij}$ which couple to the magnetic
quarks via the superpotential
\begin{equation} \label{Wmag} W\mag=M_{\rm mag} q\bar q\,.
\end{equation}
This magnetic theory flows to the same infrared fixed point. The gauge
invariant operators of both theories are in one-to-one correspondence:
\begin{eqnarray} \label{dualmap}
M &\ \lrarrow{1}\ &\mu\,M_{\rm mag}\,,\\
B &\ \lrarrow{1}\ &\sqrt{-(-\mu)^{N_c-N_f}\Lambda^{3N_c-N_f}}\:B_{\rm mag}\,.
\nonumber
\end{eqnarray}
The mass scale $\mu$ had to be introduced by dimensional analysis. The three
scales $\Lambda$, $\Lambda_{\rm mag}$ and $\mu$ are related by
\begin{equation} \label{scales}
\Lambda^{3N_c-N_f}\Lambda_{\rm mag}^{3(N_f-N_c)-N_f}=(-1)^{N_f-N_c}\mu^{N_f}\,.
\end{equation}
$N_f>3N_c$: In the infrared the theory flows to the trivial fixed point of
free quarks and gluons.
\section{Confinement from duality}
\FIGURE{$
\begin{array}{ccc}
SU(N_c),N_f\ {\rm flavors} &\quad\stackrel{\rm duality}{\lrarrow{6}}\quad
&SU(N_f-N_c),N_f\ {\rm flavors}\\
W=\mathop{\rm Tr}(mM) &&W\mag=M_{\rm mag} q\bar q+\mu\mathop{\rm Tr}(mM_{\rm mag})\\
\Bigg\downarrow &&\Bigg\downarrow\\
SU(N_c),N_f-p\ {\rm flavors} &\quad\stackrel{\rm duality}{\lrarrow{6}}\quad
&SU(N_f-N_c-p),N_f-p\ {\rm flavors}\\
\hat W=0 &&\hatW\mag=\hat M_{\rm mag} \hat q\hat{\bar q}
\end{array}
\caption{The electric theory is deformed by adding mass terms for some of the
quarks and the magnetic theory is deformed correspondingly. After having
integrated out the massive modes one finds two effective theories that are
again dual to each other. Displayed are the tree-level contributions to the
superpotentials.}
$}
Let us consider deformations \cite{Seib} of the theory described in the
previous section by mass terms $W=\mathop{\rm Tr}(mM)$, where $m$ is an $(N_f\times N_f)$-%
matrix of rank $p$. By the duality mapping (\ref{dualmap}) this corresponds to
adding a term $\mu\mathop{\rm Tr}(mM_{\rm mag})$ to the superpotential (\ref{Wmag}) in the
magnetic theory (cf figure 1). As our treatment of SQCD is restricted to
the (Wilsonian) low-energy effective action we have to integrate out the
massive modes from the deformed model. In the electric theory this just leads
to a reduction of the number of quark flavors by $p$. Therefore the low-energy
theory is an $SU(N_c)$ gauge theory with $N_f-p$ quark flavors and vanishing
superpotential $\hat W=0$. In the magnetic theory, integrating out the massive
components of $M_{\rm mag}$ leads to non\discretionary{-}{}{-}van\-ish\-ing\ expectation values for $q$, $\bar q$
and thus the gauge symmetry is broken spontaneously. The low-energy theory is
an $SU(N_f-N_c-p)$ gauge theory with $N_f-p$ quark flavors and superpotential
$\hatW\mag=\hat M_{\rm mag}\hat q\hat{\bar q}$, where hats denote the low-energy
fields. One finds that the two effective theories are again dual to each other
\cite{Seib}, as shown in figure 1.
It is interesting to consider the special case $p=N_f-N_c-1$. Then one has an
effective $SU(N_c)$ gauge theory with $N_c+1$ quark flavors on the electric
side. The magnetic gauge symmetry is completely broken by the Higgs effect.
However, one color component of each of the $N_c+1$ quark flavors stays
massless after the symmetry breaking. These $2(N_c+1)$ gauge singlets are
denoted by $\hat q$, $\hat{\bar q}$. They couple to the meson singlets
$\hat M_{\rm mag}$ via the tree-level superpotential (\ref{Wmag}). Due to
non\discretionary{-}{}{-}re\-nor\-ma\-li\-za\-tion\ theorems this is not corrected in perturbation theory. But there are
non\discretionary{-}{}{-}per\-tur\-ba\-tive\ corrections generated by instantons. The full superpotential of the
low-energy magnetic theory reads \cite{Seib}
\begin{equation} \label{Wmaglow}
\hatW\mag=\hat M_{\rm mag}\hat q\hat{\bar q}+\Lambda_{\rm mag}^{3-N_f}\det\hat M_{\rm mag}.
\end{equation}
From the fact that all physical degrees of freedom of the effective magnetic
theory are gauge invariant one expects that, as a consequence of the duality
mapping, the degrees of freedom of the electric theory are gauge singlets as
well. We will see that this intuition is right. The mapping (\ref{dualmap})
gives
\begin{eqnarray} \label{dualmaplow}
M &\ \lrarrow{1}\ &\mu\,M_{\rm mag}\,,\\
B &\ \lrarrow{1}\ &\sqrt{\mu^{-1}\Lambda^{2N_f-3}}\:\hat q\,,\nonumber
\end{eqnarray}
and the scale matching (\ref{scales}) now reads
\begin{equation} \label{scaleslow}
\Lambda^{2N_f-3}\Lambda_{\rm mag}^{3-N_f}=-\mu^{N_f}\,.
\end{equation}
The effective electric theory is described by the mesons $M$ and the baryons
$B$, $\bar B$. One can check that the 't Hooft anomaly matching conditions
\cite{tH} are satisfied for this low-energy spectrum. These conditions require
that if one gauges the global symmetries of the theory then the values of
the various triangle anomalies (which in general will not vanish) must
coincide for the microscopic description in terms of quarks and gluons and
the macroscopic description in terms of mesons and baryons. Performing this
calculation one finds that the mesons and baryons obtained from the duality
mapping (\ref{dualmaplow}) are just the right degrees of freedom to match
the global anomalies of the microscopic theory. This means that the theory is
in the confining phase \cite{Seib_D49}, in agreement with the result for
$N_f=N_c+1$ of the previous section. It is easy to determine the full
superpotential of the effective electric theory by applying the duality
mapping (\ref{dualmaplow}) to the magnetic superpotential (\ref{Wmaglow}).
Using the scale relation (\ref{scaleslow}) one finds
\begin{equation} \hat W={\bar BMB-\det M\over\Lambda^{2N_f-3}}\,,
\end{equation}
which coincides with the superpotential (\ref{Wconf}) found in the previous
section for $N_f=N_c+1$.
\section{Generalizations}
The results on SQCD described in the previous sections have been generalized
to other gauge theory models involving different gauge groups and/or
matter fields transforming in representations other than the fundamental.
For each of these models the electric theory shows confinement without
breaking of the chiral symmetry when the gauge symmetry of its magnetic dual
is completely broken.
\subsection{other gauge groups}
The simplest extension of the results on SQCD described above consists in
gauge theories with orthogonal or symplectic gauge groups.
Let us first consider an $SO(N_c)$
gauge theory with $N_f$ matter fields $Q$ (quarks) transforming in the vector
representation of the gauge group and vanishing tree-level superpotential.
This has a dual description \cite{Seib,SOdual} in terms of a magnetic
$SO(N_f+4-N_c)$ gauge theory with $N_f$ quarks $q$ transforming in the
vector representation and ${\textstyle\hf} N_f(N_f+1)$ meson singlets $M_{\rm mag}$ that
couple to the magnetic quarks via $W\mag=M_{\rm mag} qq$. The gauge invariant
operators of both theories are in one-to-one correspondence to each other by a
mapping very similar to (\ref{dualmap}). For $N_f=N_c-3$ the magnetic theory
is completely higgsed, and one finds that the electric theory confines. The
confining spectrum (mesons $M$ and baryons $B$) as well as the correct
confining superpotential
\begin{equation} W={MBB\over\Lambda^{2N_f+3}} \end{equation}
can be obtained from the effective magnetic theory via the duality mapping.
An $Sp(2N_c)$ gauge theory with $2N_f$ quarks $Q$ transforming in the
fundamental representation and vanishing tree-level superpotential can
equivalently be described \cite{IP} by an $Sp(2(N_f-2-N_c))$ gauge
theory with $2N_f$ quarks $q$ and $N_f(2N_f-1)$ meson singlets $M_{\rm mag}$ that
couple to the magnetic quarks via $W\mag=M_{\rm mag} qq$. The duality mapping
between the gauge invariant operators of the two theories is simply given
by $M\:\leftrightarrow\:\mu\,M_{\rm mag}$; there are no baryons in symplectic gauge theories.
For $N_f=N_c+2$ the magnetic theory is completely higgsed, and one finds that
the electric theory confines. The confining spectrum (mesons $M$) can be
obtained from the effective magnetic theory via the duality mapping.
To obtain the correct confining superpotential
\begin{equation} W={\mathop{\rm Pf} M\over\Lambda^{2N_f-3}} \end{equation}
more care is needed, because it is due to instanton corrections in the
effective magnetic gauge theory.
\subsection{gauge theories containing tensor fields}
For any $N=1$ supersymmetric model with vanishing tree-level superpotential
the form of the most general superpotential that can possibly be generated by
non\discretionary{-}{}{-}per\-tur\-ba\-tive\ effects is com\-plete\-ly fixed by the requirement that it be
invariant under all symmetries of the considered model \cite{ADS,ILS,confine}.
For a theory with gauge group $G$ and chiral matter fields $\phi_l$ in
representations $r_l$ of $G$ and dynamically generated scale $\Lambda$
one finds
\begin{equation} \label{Weff}
W \propto \left({\prod_l(\phi_l)^{\mu_l}\over\Lambda^b}\right)^{2\over\Delta}
\,,
\end{equation}
where $\mu_l$ is the (quadratic) Dynkin index of the representation $r_l$,
$\mu_G$ denotes the index of the adjoint representation,
$\Delta=\sum_l\mu_l-\mu_G$ and $b={\textstyle\hf}(3\mu_G-\sum_l\mu_l)$ is the coefficient
of the 1-loop $\beta$-function. In general the complete non\discretionary{-}{}{-}per\-tur\-ba\-tive\
superpotential consists of a sum of terms of the form (\ref{Weff})
with different possible contractions of all gauge and flavor indices.
The relative coefficients of these terms cannot be fixed by symmetry
arguments but must be inferred from a different reasoning.
If the theory is confining at every point of the moduli space then the
superpotential must either vanish or be a smooth function of the confined
degrees of freedom \cite{confine}. All such models with a smooth confining
superpotential could be classified \cite{confine} as they have to verify the
constraint $\Delta=2$. But only for some of these smoothly confining gauge
theories containing tensor fields a dual description in terms of magnetic
variables is known.
On the other hand many dualities for models including tensor fields have been
found once an appropriate tree-level superpotential for the tensors is added.
Let us review one example first studied by the authors of \cite{KuSch}. They
considered an $SU(N_c)$ gauge theory with $N_f$ quark flavors $Q$, $\bar Q$, an
additional matter field $X$ in the adjoint representation and a tree-level
superpotential $W_{\rm tree}=\mathop{\rm Tr} X^{k+1}$, where $k>1$ is some integer. This model
has a dual description in terms of an $SU(kN_f-N_c)$ gauge theory with $N_f$
quark flavors $q$, $\bar q$, an adjoint tensor $Y$, $kN_f^2$ singlets
$M_{{\rm mag},j}$, $j=0,\ldots,k-1$ and tree-level superpotential
\begin{equation}
W\mag=\mathop{\rm Tr} Y^{k+1}\:+\:\sum_{j=0}^{k-1}M_{{\rm mag},k-1-j}qY^j\bar q\,.
\end{equation}
For $N_c=kN_f-1$ the magnetic theory is completely higgsed and one expects the
electric theory to confine. Indeed, one finds that the 't Hooft anomaly
matching conditions are satisfied if the confined spectrum of the electric
theory is given by \cite{KSS}
\begin{eqnarray} \label{confspec}
M_j &= &QX^j\bar Q\,,\quad j=0,\ldots,k-1\,,\\
B &= &(Q)^{N_f}\cdots(X^{k-1}Q)^{N_f}(X^kQ)^{N_f-1}\,,\nonumber\\
\bar B &= &(\bar Q)^{N_f}\cdots(X^{k-1}\bar Q)^{N_f}(X^k\bar Q)^{N_f-1}\,,
\nonumber
\end{eqnarray}
where the gauge indices are contracted with a Kronecker delta for the mesons
and with an epsilon tensor of rank $N_c$ for the baryons. In addition the
flavor indices of the baryons are contracted with an epsilon tensor of rank
$kN_f$ leaving $2N_f$ independent baryons. It is easy to see \cite{mklein}
that these are exactly the degrees of freedom that are mapped under duality
on the magnetic singlets $\hat M_{{\rm mag},j}$, $\hat q$, $\hat{\bar q}$ that
stay massless after the Higgs effect. The matching of the 't Hooft anomalies
between the microscopic (i.e.\ quarks and gluons) and the macroscopic (i.e.\
confined) description of the electric theory can thus be seen as a consequence
of the anomaly matching between the electric and the magnetic theory. In this
sense confinement can be derived from duality.
\TABLE{
$\begin{array}{|c|c|c|c|c|} \hline
\rule{0mm}{3ex} &\multicolumn{4}{c|}{SU(N_c)}\\ \hline
\rule{0mm}{3ex} {\rm tensors} &adj &\Yasym+\overline{\Yasym} &\Ysym+\overline{\Ysym}
&\Yasym+\overline{\Ysym}\\
\rule{0mm}{3ex} W_{\rm tree} &X^{k+1} &(X\bar X)^{k+1} &(X\bar X)^{k+1}
&(X\bar X)^{2(k+1)}\\
\rule{0mm}{3ex} N_c &kN_f-1 &(2k+1)N_f-4k-1 &(2k+1)N_f+4k-1 &(4k+3)(N_f+4)-1\\
\rule{0mm}{3ex} \alpha &k &1 &2(k+1) &2(k+1) \\
\rule{0mm}{3ex} \beta_X &(k-1)N_c &k(N_f-1) &(k+1)(2k(N_f+2)-1)
&\msmall{2(k+1)((2k+1)(N_f+4)-3)}\\
\rule{0mm}{3ex} \beta_{\bar X} & &k(N_f-1) &(k+1)(2k(N_f+2)-1)
&\msmall{2(k+1)((2k+1)(N_f+4)+1)}\\
\rule{0mm}{3ex} \gamma &-N_c &3-N_f &-(N_c+N_f+4) &-2(k+1)(N_f+4)\\ \hline
\end{array}$\phantom{xx}
$\begin{array}{|c|c|c|c|c|} \hline \rule{0mm}{3ex} &\multicolumn{2}{c}{Sp(2N_c)}
&\multicolumn{2}{|c|}{SO(N_c)}\\ \hline
\rule{0mm}{3ex} {\rm tensors} &\Ysym &\Yasym &\Yasym &\Ysym\\
\rule{0mm}{3ex} W_{\rm tree} &X^{2(k+1)} &X^{k+1} &X^{2(k+1)} &X^{k+1}\\
\rule{0mm}{3ex} N_c &(2k+1)N_f-2 &k(N_f-2) &(2k+1)N_f+3 &k(N_f+4)-1\\
\rule{0mm}{3ex} \alpha &2k+1 &1 &1 &k\\
\rule{0mm}{3ex} \beta &2k(N_c+1) &(k-1)(N_f-1) &2k(N_f+1)+4 &(k-1)(N_c-2k)\\
\rule{0mm}{3ex} \gamma &-(N_c+1) &3-N_f &1-N_f &-(N_c+2k)\\ \hline
\end{array}$
\caption{Gauge theories that confine in the presence of a
tree-level superpotential. The microscopic spectrum consists of $N_f$
quark flavors and additional fields transforming in tensor representations
represented by their Young tableaux. The coefficients $\alpha$, $\beta$,
$\gamma$ refer to the powers in the non\discretionary{-}{}{-}per\-tur\-ba\-tive\ superpotential (\ref{Weff2}).}
}
\TABLE{
$\begin{array}{|c|c|c|c|c|} \hline
\rule{0mm}{3ex} &\multicolumn{4}{c|}{SU(N_c)}\\ \hline
\rule{0mm}{3ex} {\rm tensors} &2\,adj &adj+\Yasym+\overline{\Yasym} &adj+\Ysym+\overline{\Ysym}
&adj+\Yasym+\overline{\Ysym}\\
\rule{0mm}{3ex} W_{\rm tree} &X^{k+1}+XY^2 &X^{k+1}+XY\bar Y &X^{k+1}+XY\bar Y
&X^{k+1}+XY\bar Y\\
\rule{0mm}{3ex} N_c &3kN_f-1 &3kN_f-5 &3kN_f+3 &3k(N_f+4)-1\\
\rule{0mm}{3ex} \alpha &3k &k &4k &2k \\ \hline
\end{array}$\phantom{xx}
$\begin{array}{|c|c|c|c|c|} \hline \rule{0mm}{3ex} &\multicolumn{2}{c}{Sp(2N_c)}
&\multicolumn{2}{|c|}{SO(N_c)}\\ \hline
\rule{0mm}{3ex} {\rm tensors} &2\,\Yasym &\Yasym+\Ysym &2\,\Ysym
&\Ysym+\Yasym\\
\rule{0mm}{3ex} W_{\rm tree} &X^{k+1}+XY^2 &X^{k+1}+XY^2 &X^{k+1}+XY^2
&X^{k+1}+XY^2\\
\rule{0mm}{3ex} N_c &3kN_f-4k-2 &3kN_f-4k+2 &3kN_f+8k+3 &3kN_f+8k-5\\
\rule{0mm}{3ex} \alpha &(k) &(3k) &3k &k\\ \hline
\end{array}$
\caption{Gauge theories that confine in the presence of a
tree-level superpotential. The microscopic spectrum consists of $N_f$
quark flavors and additional fields transforming in tensor representations
represented by their Young tableaux. The coefficient $\alpha$ refers to the
power in the non\discretionary{-}{}{-}per\-tur\-ba\-tive\ superpotential (\ref{Weff2}).}
}
It is straightforward to apply this idea to all gauge theory models of
\cite{ILSdual,patterns} based on simple gauge groups. One first builds the
gauge invariant composite operators whose expectation values span the moduli
space, then finds the duality mapping between the electric and the magnetic
theory for these operators and finally applies this mapping to the completely
higgsed effective magnetic theory to obtain the confined spectrum of the
electric theory. In addition, in most cases at least some of the terms of the
confining superpotential can be determined from the magnetic tree-level
superpotential. To constrain the possible form of the confining superpotential
we would like to generalize the formula (\ref{Weff}) to the case of a non\discretionary{-}{}{-}van\-ish\-ing\
tree-level superpotential. Therefore divide the matter fields into two subsets
$\{\phi_l\}=\{\bar\phi_{\bar l}\}\cup\{\hat\phi_{\hat l}\}$,
with $\{\bar\phi_{\bar l}\}\cap\{\hat\phi_{\hat l}\}=\emptyset$,
and add a tree-level term for the hatted fields:
\begin{equation} \label{Wtreehat}
W_{\rm tree}=h\,\prod_{\hat l}\left(\hat\phi_{\hat l}\right)^{n_{\hat l}},
\end{equation}
where $h$ is a dimensionful coupling parameter and the $n_{\hat l}$
are positive integers. To be invariant under all global symmetries the full
superpotential must be of the form \cite{mklein}
\begin{equation} \label{Weff2}
W \propto\left({\prod_{\bar l}(\bar\phi_{\bar l})^{\mu_{\bar l}}\over\Lambda^b}
\right)^\alpha\ \prod_{\hat l}\left(\hat\phi_{\hat l}
\right)^{\beta_{\hat l}}\ h^\gamma\,,
\end{equation}
where the powers $\alpha$, $\beta_{\hat l}$, $\gamma$ must verify the
following relations:
\begin{eqnarray} \label{albega}
\gamma &= &1-{\textstyle\hf}\alpha\Delta\,,\nonumber \\
\beta_{\hat l} &= &\mu_{\hat l}\,\alpha+\gamma n_{\hat l}\,.
\end{eqnarray}
The calculation of the confining spectra and the confining superpotentials
for all simple group models of \cite{ILSdual,patterns} is performed in
\cite{mklein}. The results are displayed in tables 1 and 2.
\section{A new confining model}
To illustrate the ideas presented above I would like to treat one of the models
of \cite{mklein} in more detail. Consider an $SU(N_c)$ gauge theory with
$N_f+8$ quarks $Q$, $N_f$ antiquarks $\bar Q$, an antisymmetric\ tensor $X$ and a
conjugate symmetric tensor $\bar X$ and tree-level superpotential
$W_{\rm tree}=h\,\mathop{\rm Tr}(X\bar X)^{2(k+1)}$. This is a chiral theory and the difference
between the number of quarks and antiquarks is required by anomaly freedom.
The gauge invariant composite operators are given by
\begin{eqnarray} \label{mesbarsas}
&& M_j=Q\bar Q_{(j)}\,,\quad
P_r= Q\bar XQ_{(r)}\,,\ \bar P_r=\bar QX\bar Q_{(r)}\,,\nonumber\\
&&\quad{\rm with\ } Q_{(j)}=(X\bar X)^j Q\,,\ \bar Q_{(j)}=(\bar XX)^j\bar Q\,,
\nonumber\\
&& \quad j=0,\ldots,2k+1\,,\quad r=0,\ldots,2k\,,\nonumber \\
&& \bar{\cal B}^{(\bar n_0,\ldots,\bar n_{2k},n_0,\ldots,n_{2k+1})}
=(\bar X(X\bar X)^kW_\alpha)^2\nonumber\\
&& \qquad \cdot(\bar XQ)^{\bar n_0}(\bar XQ_{(1)})^{\bar n_1}
\cdots(\bar XQ_{(2k)})^{\bar n_{2k}}\nonumber\\
&& \qquad \cdot\bar Q^{n_0}\bar Q_{(1)}^{n_1}\cdots
\bar Q_{(2k+1)}^{n_{2k+1}}\,,\nonumber\\
&& \quad{\rm with\ } \sum_{j=0}^{2k+1}n_j+\sum_{j=0}^{2k}\bar n_j=N_c-4\,,\\
&& B_n=X^n Q^{N_c-2n}\,,\quad
n=0,\ldots,\left\lceil{N_c\over2}\right\rceil\,,\nonumber\\
&& \bar B_{\bar n}=\bar X^{\bar n} \bar Q^{N_c-\bar n}
\bar Q^{N_c-\bar n}\,,\quad \bar n=0,\ldots,N_c\,,\nonumber\\
&& T_i=\mathop{\rm Tr}(X\bar X)^i\,,\quad i=1,\ldots,2k+1\,,\nonumber
\end{eqnarray}
where the gauge indices are contracted with one epsilon tensor for the
$\bar{\cal B}^{(\cdots)}$, $B_n$ and with two epsilon tensors for the
$\bar B_{\bar n}$.
The authors of \cite{ILSdual} found a dual description of this model in terms
of a magnetic $$SU((4k+3)(N_f+4)-N_c)$$ gauge theory, with $N_f+8$ quarks $q$,
$N_f$ antiquarks $\bar q$, an antisymmetric\ tensor $Y$, a conjugate symmetric tensor
$\bar Y$ and singlets $M_{{\rm mag},j}$, $P_{{\rm mag},r}$,
$\bar P_{{\rm mag},r}$ and tree-level superpotential
$$W\mag=-h\,\mathop{\rm Tr}(X\bar X)^{2(k+1)}+\ldots\,,$$ where the dots indicate terms
involving $M_{\rm mag}$, $P_{\rm mag}$, $\bar P_{\rm mag}$. The duality mapping for the gauge
invariant operators is given by
$$M_j,\,P_r,\,\bar P_r\ \leftrightarrow\
M_{{\rm mag},j},\,P_{{\rm mag},r},\,\bar P_{{\rm mag},r}\,,$$
\begin{eqnarray} \label{barmapsas}
\bar{\cal B}^{(\bar n_i,n_j)} &\leftrightarrow &\bar{\cal B}_{\rm mag}^{(\bar m_i,m_j)}\,,\nonumber\\
{\rm with\ } && m_j=N_f-n_{2k+1-j}\,,\nonumber\\
&&\bar m_j=N_f+8-\bar n_{2k-j}\,,\\
B_n &\leftrightarrow &B_{{\rm mag},m}\,,\nonumber\\\
{\rm with\ } && m=(2k+1)(N_f+4)-2-n\,,\nonumber\\
\bar B_{\bar n} &\leftrightarrow &\bar B_{{\rm mag},\bar m}\,,\nonumber\\
{\rm with\ } && \bar m=2(2k+1)(N_f+4)+4-\bar n\,.\nonumber
\end{eqnarray}
The dependence on the mass scale $\mu$ has been suppressed.
For $N_c=(4k+3)(N_f+4)-1$ the magnetic theory is completely higgsed and the
electric theory confines with low-energy spectrum given by the composite
fields
\begin{eqnarray} \label{confspecsas}
&&M_j\,,\ P_r\,,\ \bar P_r\,,\nonumber\\
&&\quad j=0,\ldots,2k+1\,,\quad r=0,\ldots,2k\,,\nonumber\\
&&B\equiv B_{(2k+1)(N_f+4)-2}\,,\nonumber\\
&&\bar B\equiv\bar{\cal B}^{(N_f+8,\ldots,N_f+8,N_f,\ldots,N_f,N_f-1)}\,,\\
&&\bar b\equiv \bar B_{2(2k+1)(N_f+4)+3}\,,\nonumber
\end{eqnarray}
of eqs.\ (\ref{mesbarsas}). From (\ref{barmapsas}) one finds the mappings
$B,\bar B\leftrightarrow q,\bar q$ and $\bar b\leftrightarrow \bar Y$. One color component of
each of the fields $q$, $\bar q$, $\bar Y$ together with the meson singlets
are exactly the degrees of freedom that stay massless after breaking the
magnetic gauge group.
As a further consistency check let us consider deformations of the theory
along the flat directions corresponding to large expectation values of the
baryons $B$, $\bar b$. A large VEV of $B$ breaks the gauge symmetry to
$Sp(2((2k+1)(N_f+4)-2))$ \cite{ILSdual}. The low-energy theory contains
$2(N_f+4)$ quarks $Q$, a symmetric tensor $X$ and tree-level superpotential
$\mathop{\rm Tr} X^{2(k+1)}$. This model is known to show confinement \cite{newconf}.
A large VEV of $\bar b$ breaks the gauge symmetry to $SO(2(2k+1)(N_f+4)+3)$
\cite{ILSdual}. The low-energy theory contains $2(N_f+4)$ quarks $Q$, an antisymmetric\
tensor $X$ and tree-level superpotential $\mathop{\rm Tr} X^{2(k+1)}$. This model is known
to show confinement \cite{newconf}.
The effective low-energy superpotential of the magnetic theory contains the
terms $M_{2k+1}q\bar q\ +\ P_{2k}q\bar Yq$.
We thus expect that the confining superpotential of the electric theory has
terms proportional to $\bar B M_{2k+1} B$, $BBP_{2k}\bar b$. The detailed
analysis gives \cite{mklein}
\begin{eqnarray} \label{Wconfsas}
W &= &{\bar B M_{2k+1} B\over h^{2(k+1)(N_f+4)}\,
\Lambda^{2(k+1)((8k+5)(N_f+4)-2)}}\nonumber\\
&&\ +\ {BBP_{2k}\bar b\over h^{2(N_f+4)-1}\Lambda^{2((8k+5)(N_f+4)-2)}}\\
&&\ +\ \ldots\,,\nonumber
\end{eqnarray}
where the dots stand for possible further terms that could be generated by
instanton effects in the completely broken magnetic gauge group.
\section{Conclusion}
I have shown how the non\discretionary{-}{}{-}Abel\-ian\ duality of $N=1$ supersymmetric gauge theories
discovered by Seiberg can be used to find new models that confine in the
presence of an appropriate superpotential.
This is a very interesting application of the proposed duality because
it enables us to obtain non\discretionary{-}{}{-}per\-tur\-ba\-tive\ results for the electric theory by
a perturbative calculation in its magnetic dual. Confinement in the
electric theory can be understood from the Higgs phase of the magnetic
theory. The confining spectrum can easily be derived from the
duality mappings of gauge invariant operators. For $SU$ and $SO$
gauge groups one also obtains the form of the confining superpotential
by applying these mappings to the magnetic tree-level superpotential.
To determine the full confining superpotential one needs to include
instanton corrections in the completely broken magnetic gauge group.
For $Sp$ gauge groups the tree-level superpotential of the completely
higgsed magnetic theory vanishes. In this case the whole magnetic
superpotential is non\discretionary{-}{}{-}per\-tur\-ba\-tive\ and therefore more difficult to obtain.
\acknowledgments
I would like to thank the organizers for this very interesting workshop
and for their warm hospitality. I would also like to thank L.~Ib\'a\~nez,
\linebreak
C.~Cs\'aki and H.~Murayama for helpful discussions and e-mail correspondence.
This work is supported by the European Union under grant FMRX-CT96-0090.
|
\section{Introduction}
In recent years, supersymmetry (SUSY) has become increasingly favoured
as the theoretical structure underlying fundamental particle
interactions. In this light it is natural to investigate possible
cosmological implications of SUSY.
Until recently, inflation and topological defects were thought of as
rival theories for the formation of density perturbations in the
Universe. However, in supersymmetric theories it is natural for the
two to coexist. In this paper we examine recently proposed
supersymmetric models which give rise to cosmic strings
\cite{nonabelian,abelian}. We show that these theories are
automatically hybrid inflation models. Previous work explored the
microphysics of the cosmic strings, but did not consider the resulting
cosmology. Whilst the cosmology of mixed models with cosmic strings
and inflation has previously been explored, the fact that
supersymmetric models with cosmic strings naturally inflate is not so
well known.
The theories we examine are based on both non-Abelian and Abelian
gauge symmetries, with symmetry breaking arising in two distinct ways.
The first model is a $SU(2) \times U(1) \rightarrow U(1) \times Z_2$
supersymmetric gauge theory with the gauge symmetry breaking
implemented with a $F$ term\cite{nonabelian}. The second theory is a
SUSY $U(1) \rightarrow Z_2$ gauge theory with a
Fayet-Illiopoulos $D$ term\cite{abelian}. In both cases it was
previously found that there were fermion zero modes and supersymmetry
was broken in the string core. The zero modes in the first theory
occurred in pairs, whilst there was a single zero mode in the second
theory. The effects of soft-SUSY breaking terms were investigated. In
the former case: (1) SUSY breaking Higgs mass terms did not destroy
the zero modes; (2) SUSY breaking gaugino mass terms destroyed all
zero modes which involved gauginos; (3) SUSY violating trilinear terms
created Yukawa couplings which destroyed all zero modes. In the
second theory, the zero mode survived soft-SUSY breaking terms. This
zero mode could move along the string resulting in the string carrying
a current. Such current-carrying strings could cause cosmological
problems \cite{vortons}.
The models we consider are particularly attractive for several
reasons. (1) They incorporate both cosmic strings and hybrid
inflation, which has recently been shown to give good agreement with
observations \cite{joao,battye}. (2) The inflaton field is typically much
smaller than $M_P$, allowing Planck era physics to be ignored. (3)
Inflaton physics can be taken to be normal GUT particle
physics. Normally the inflaton potential must suddenly ``turn off'' to
allow inflation to end. For field values less than $M_P$, such
functions are often complicated. In hybrid inflation, the ending of
inflation is simply done by one of the other GUT fields. This
simplification allows many types of simple potentials to cause and end
inflation. (4) By invoking supersymmetry, coupling constants often do
not need to be fine-tuned. \cite{linde,copeland,lyth}
In this paper we show that the two models we consider do indeed hybrid
inflate and lead to acceptable cosmic microwave
anisotropies. Supersymmetry is broken during inflation and a radiative
correction to the effective potential ensures that inflation ends.
SUSY is restored after cosmic strings are formed, though as previously
shown \cite{abelian}, it is broken in the string core. Inflation ends
before the GUT phase transition commences since the slow roll
condition is violated before the phase transition. Similar results
have appeared in different contexts
\cite{lyth,dvali1,dvali3,jean1,jean2}.
We proceed by reviewing the $SU(2) \times U(1)$ $F$-term SUSY gauge
theory and the $D$-term U(1) model. Then we compute the radiative
correction to the effective potential for both models. The CMB
anisotropy is then calculated. Finally we end with our conclusions.
\section{A $SU(2) \times U(1)$ F-term SUSY gauge theory}
The tree-level effective potential for a SUSY invariant theory is
\begin{equation}
U = \sum_i |F_i|^2 + \frac{1}{2}
\sum_a D^a D_a
\label{susypot}
\end{equation}
\noindent
where the $F$-term (first term) is $F_i = \delta W[\phi_i] /\delta
\phi_i$, and $W[\phi_i]$ is the superpotential with the fermionic
terms set to zero. The $D$ term (second term) is $D^a = -(\kappa +
g\sum_{ij} q_a \phi_i^{\dagger} (T^a)^i_j \phi^j $. Here $\kappa$ is a
Fayet-Illiopoulos constant and is nonzero if and only if the action
possesses a $U(1)$ symmetry. $T^a$ is the group generator and the sum
in (\ref{susypot}) is over all group generators. $g$ is the coupling
constant and $q_a$ is a dimensionless charge.
The $SU(2) \times U(1) $ theory we are considering has the
superpotential\cite{nonabelian}
\begin{equation}
W = \mu_1 S_0 ({\bf{\Phi}} \cdot {\bf{\tilde{\Phi}}} - \eta^2) + \mu_2 (S
{\bf{\Phi}}^T \Lambda {\bf{\Phi}} +
\tilde{S} {\bf{\tilde{\Phi}}}^T \Lambda {\bf{\tilde{\Phi}}})
\end{equation}
\noindent
where ${\bf{\Phi}}$ and ${\bf{\tilde{\Phi}}}$ transform as $SU(2)$
triplets, and $S_0, S, \tilde{S}$ are $U(1)$ scalars.
The tree level effective potential is
\begin{eqnarray}
U & = & \mu_1^2 | {\bf{\phi}} \cdot {\bf{\tilde{\phi}}} -\eta^2|^2 + \mu_2^2
|2\phi_1 \phi_3 - \phi_2^2|^2 + \mu_2^2 | 2 \tilde{\phi_1}
\tilde{\phi_3} - \tilde{\phi_2}^2|^2 + \nonumber\\
& {} & |\mu_1 s_0 {\bf{\tilde{\phi}}} +2 \mu_2 s \Lambda \phi |^2 +
|\mu_1 s_0 {\bf{\phi}} + 2 \mu_2 s \Lambda \tilde{\phi} |^2 + \nonumber\\
&{}& e^2 | (\phi_1 +\phi_3) \phi^*_2 -
(\tilde{\phi_1} + \tilde{\phi_3}) \tilde{\phi}_2^*|^2 + \nonumber\\
& {} & \frac{e^2}{2}
(|\phi_1|^2 -|\phi_3|^2 -|\tilde{\phi_1}|^2 -|\tilde{\phi_3}|^2)^2 + \nonumber\\
& {} & \frac{e^2}{3} (|{\bf{\phi}}|^2 -|{\bf{\tilde{\phi}}}|^2 -2|s|^2
-2|\tilde{s}|^2 )^2. \nonumber
\end{eqnarray}
The minimum of the potential with regard to variations in the $SU(2)$
fields, $\phi, \tilde{\phi}$, leads to the constraint $\phi_1 =
\tilde{\phi_1}$ and $\phi_2 = \tilde{\phi_2} = \phi_3 = \tilde{\phi_3}
= 0$. The potential then becomes
\begin{equation}
U = \mu_1^2 | \phi_1^2 -\eta^2|^2 + 2 \mu_1^2
|s_0|^2 | {\phi_1}|^2.
\end{equation}
\noindent
This possesses two minima. A global minimum exists at $\phi_1 = \eta \equiv
M_{s_0}, s_0 = 0$ where supersymmetry is unbroken. There is also a
SUSY violating local minimum at $\phi_1=0, s_0 > \eta \equiv
s_0^{crit}$. This corresponds to the false vacuum. We will assume a
chaotic inflation scenario, where $s_0$ is initially a random field
and $s_0 > s_0^{crit}$. Thus the $SU(2)$ fields are initially trapped
in a false vacuum and the potential initially is
\begin{equation}
U = \mu_1^2 \eta^4.
\end{equation}
\noindent
The potential has no slope prompting $s_0$ to roll down to its
minimum. However, eternal inflation is averted since SUSY is broken
and the tree level potential is no longer protected from radiative
corrections which cause inflation to end.
\section{A $U(1)$ $D$-term Gauge Theory}
In this $U(1)$ theory we assume there exists a nonzero Fayet-Illiopoulos
constant, $\kappa$. The superpotential is taken to be \cite{abelian}
\begin{equation}
W = \alpha \Phi_0 \Phi_+ \Phi_-
\end{equation}
\noindent
where $\Phi_0, \Phi_+, \Phi_-$ have $U(1)$ charges 0, +1, -1,
respectively.
If we redefine $D^a = -g(\kappa + \sum_{ij} q_a \phi_i^{\dagger}
(T^a)^i_j \phi^j) $, as is often done, the tree-level potential is \cite{lyth,jean2,dvali2}
\begin{eqnarray}
U & = & \alpha^2|\phi_0|^2(|\phi_+|^2 + |\phi_-|^2) + \alpha^2 |\phi_+
\phi_-|^2 \nonumber\\
& {} &
+ \frac{e^2}{2}(|\phi_+|^2 -|\phi_-|^2 + \kappa)^2.
\end{eqnarray}
\noindent
$U$ possesses two minima -- a global minimum at $\phi_0 = \phi_+ =0,
\phi_- = \sqrt{\kappa}$ where supersymmetry is unbroken, and a SUSY
violating local minimum at $\phi_+ = \phi_- = 0, \phi_0 > (e/\alpha)
\sqrt{\kappa} = \phi_0^{crit}$. As before, we assume a chaotic
inflation scenario and that $\phi_0 > \phi_0^{crit}$. The fields are
in the false vacuum and the potential is
\begin{equation}
U = \frac{e^2}{2} \kappa^2.
\end{equation}
\noindent
As before this potential has no slope. However, a SUSY breaking
induced radiative correction will add a slope.
\section{One Loop Radiative Corrections to the $SU(2) \times U(1)$ and $U(1)$
Theories} \label{correction}
\subsection{ $SU(2) \times U(1)$ Theory}
We first consider the one loop radiative correction for the $SU(2)
\times U(1)$ theory after SUSY breaking.
At one loop, $\phi$ and $\tilde{\phi}$ receive mass corrections. The
$s, \tilde{s}, s_0$ fields will not acquire masses. To see this,
shift the fields from their expectation values
\begin{eqnarray}
s \rightarrow s' + s \,\,\,\,\,
& \tilde{s} \rightarrow \tilde{s}' + \tilde{s} &
\,\,\,\,\, s_0 \rightarrow s_0' + s_0 \nonumber\\
\phi_i \rightarrow \phi_i' + \phi_i & {} &
\tilde{\phi}_i \rightarrow \tilde{\phi}_i' + \tilde{\phi}_i
\end{eqnarray}
\noindent
where we take $s,\tilde{s},s_0,\phi_i,\tilde{\phi}_i$ to be the classical
vacuum expectation values. We assume that we are in the false vacuum
dominated case. Thus
\begin{eqnarray}
s = \tilde{s} = \phi_i = \tilde{\phi}_i = 0\,\,\, &{}&\,\,\,
s_0 \neq 0
\end{eqnarray}
Upon substituting the shifted fields, and requiring $\phi_i' =
\tilde{\phi_i'}$ and $s = \tilde{s}$, the $D$-terms drop out. The
potential to leading order is
\begin{equation}
U = \mu_1^2 | \vec{\phi}'^2 -\eta^2|^2 + 2 \mu_1^2 |s_0|^2 |
{\vec{\phi}}'|^2 + \dots
\end{equation}
\noindent
To find the masses of the fields substitute
\begin{equation}
\phi_i' = \frac{a_i + ib_i}{\sqrt{2}}.
\end{equation}
\noindent
This gives
\begin{eqnarray}
U & = & 2 \cdot \frac{(\mu_1^2 |s_0|^2 - \mu_1^2 \eta^2)}{2} \sum_i
a_i^2 \nonumber \\
& {} & 2 \cdot \frac{( \mu_1^2 |s_0|^2 + \mu_1^2
\eta^2)}{2} \sum_j b_j^2 + {\rm non-mass\,\, terms}. \nonumber\\
\end{eqnarray}
Thus the $a_i$ acquire masses $ \mu_1^2 |s_0|^2 - \mu_1^2 \eta^2$ and
the $b_i$ acquire masses $\mu_1^2 |s_0|^2 + \mu_1^2 \eta^2 $. (There
are two sets of $a_i$ fields and two sets of $b_i$ -- hence the factor
of 2.) The $\vec{\phi}$ ($\vec{\tilde{\phi}}$) field splits up
into two fields \cite{dvali1,dvali2,lyth}, $\vec{\phi} +
\vec{\phi^*}$,($\vec{\tilde{\phi}} + \vec{\tilde{\phi^*}}$) and
$i(\vec{\phi} - \vec{\phi^*})$, ($i(\vec{\tilde{\phi}} -
\vec{\tilde{\phi^*}})$) with masses $ \mu_1 |s_0|^2 \pm \mu_1^2
\eta^2$. We now calculate the masses of the fermion partners.
For a general superpotential
\begin{equation}
W[\Phi] = f_i \Phi_i + \frac{m_{ij}}{2} \Phi_i \Phi_j +
\frac{\lambda_{ijk}}{3} \Phi_i \Phi_j \Phi_k
\end{equation}
\noindent
the fermionic mass portion of the Lagrangian is
\begin{equation}
{\cal{L}}^F_m = -\frac{1}{2} ( m_{ij} + 2 \lambda_{ijk} \langle \phi_k
\rangle ) \psi_i
\psi_j + {\rm H.C}
\end{equation}
In our case $m_{ij} = 0$ and the only non-zero scalar expectation
value is $s_0$. Thus,
\begin{equation}
{\cal{L}}^F_m = -\mu_1 s_0 \psi_i
\tilde{\psi}_j + {\rm H.C},
\end{equation}
\noindent
and the fermions have mass $\mu_1 s_0$.
The 1-loop potential is found using the formula\cite{coleman}
\begin{equation}
U(s_0)_{1-loop} = \sum \frac{(-)^F}{64 \pi^2} M_i(s_0)^4 \ln
\frac{M_i(s_0)^2}{\Lambda^2}
\end{equation}
\noindent
where $\Lambda$ is an ultraviolet cutoff, and $F = \pm1$ for bosons/fermions.
Inserting the masses and adding in the zeroth order
part we have \cite{dvali1,lyth}
\begin{eqnarray}
U_{eff}(s_0) & = & \mu_1^2\eta^4(1 + 2 \cdot \frac{\mu_1^2}{64 \pi^2}(2
\ln(\frac{ \mu_1 |s_0|}{\Lambda})^2 \nonumber\\
&{}&
+ (\frac{{s_0}^2}{\eta^2} - 1)^2 \ln (1 - \frac{\eta^2}{{s_0}^2})
+ (\frac{{s_0}^2}{\eta^2} + 1)^2 \ln (1 + \frac{\eta^2}{{s_0}^2})) \nonumber.
\end{eqnarray}
In the chaotic inflation limit, $s_0 \gg s_0^{crit} = \eta$, one finds
\begin{equation}
U_{eff}(s_0) = \mu_1^2 \eta^4 (1 + \frac{\mu_1^2}{16 \pi^2}(\ln
(\frac{ \mu_1 |s_0|}{\Lambda})^2 + \frac{3}{2})).
\end{equation}
\subsection{U(1) Theory with Fayet-Illiopoulos $D$-term}
The one loop correction for the $U(1)$ theory with a nonzero
Fayet-Illiopoulos constant, $\kappa$, is computed just as for the
$SU(2) \times U(1)$ theory. As before, the $\phi_+$ and $\phi_-$ split
up into scalars with masses $\alpha^2 |\phi_0|^2 \pm e^2 \kappa$. The
one loop potential in the limit of a large inflaton field $\phi_0$ is
\cite{dvali2,lyth}
\begin{equation}
U_{eff}(\phi_0) = \frac{e^2 \kappa^2}{2}(1 + \frac{e^2}{16 \pi^2}(\ln
(\frac{ \alpha |\phi_0|}{\Lambda})^2 + \frac{3}{2})).
\end{equation}
\section{The End of Inflation}
Inflation occurs as long as the slow-roll parameters $\epsilon, \eta$
remain small, ($\epsilon, \eta << 1$), where \cite{copeland}
\begin{eqnarray}
\epsilon = \frac{M_P^2}{16 \pi} (\frac{U_{eff}'}{U_{eff}})^2 \,\,\,
&{}& \,\,\,
\eta = \frac{M_P^2}{8 \pi} \frac{U_{eff}''}{U_{eff}}
\end{eqnarray}
Writing $s_0 = x s_0^{crit}$ and using
\begin{equation}
U'_{eff} = \frac{ \mu_1^4 M_{s_0}^3}{32 \pi^2} 4x((x^2-1)\ln(1-\frac{1}{x^2}) +
(x^2+1)\ln(1+\frac{1}{x^2}))
\end{equation}
In the limit of large $x$, this is
\begin{equation}
U'_{eff} \approx \frac{\mu_1^4 M_{s_0}^3}{8 \pi^2 } \frac{1}{x}.
\label{uprime_lim}
\end{equation}
\noindent
We find for the $SU(2) \times U(1)$ case \cite{dvali1}
\begin{eqnarray}
\epsilon & = & (\frac{\mu_1^2 M_P}{8\pi^2 M_{s_0}})^2 \frac{x^2}{16 \pi}
((x^2-1) \ln(1 - \frac{1}{x^2}) + (x^2 + 1) \ln(1 + \frac{1}{x^2}))^2
\label{ep_limit} \nonumber\\
\eta & = & (\frac{\mu_1 M_P}{4\pi M_{s_0}})^2 \frac{1}{4 \pi}
((3x^2-1) \ln(1 - \frac{1}{x^2}) + (3 x^2 + 1) \ln(1 + \frac{1}{x^2}))
\label{eta_lim} \nonumber
\end{eqnarray}
\noindent
where we approximated $U_{eff} = \mu_1^2 \eta^4$. In the limit of
large $x$
\begin{eqnarray}
\epsilon & = & \frac{\mu_1^4}{1024 \pi^5}(\frac{M_P}{M_{s_0}})^2 \frac{1}{x^2}
\nonumber\\
\eta & = & -\frac{\mu_1^2}{64 \pi^3} (\frac{M_P}{ M_{s_0}})^2 \frac{1}{x^2}.
\end{eqnarray}
For $x \gg 1$, both $\eta, \epsilon$ are small. For $x \approx 1$, which
corresponds to the GUT phase transition, both $\epsilon,\eta$
diverge. Hence inflation ends \cite{dvali1} before the phase
transition; $x>1$.
The same conclusion holds for the $U(1)$ theory with a
Fayet-Illiopoulos $D$-term.
\section{Predicted CMB Anisotropies of the $SU(2) \times U(1)$ and
$U(1)$ theories}
\subsection{$SU(2) \times U(1)$ theory}
The temperature quadrapole anisotropy due to inflation can be found to
be \cite{densitypert}
\begin{eqnarray}
(\frac{\Delta T}{T})_{infl} & \approx & \sqrt{\frac{32 \pi}{45 M_P^6}} \frac{
U^{3/2}_{eff}}{U'_{eff}}|_{x_Q} \nonumber\\
& \approx & \sqrt{\frac{32 \pi}{45}} \frac{8 \pi^2}{\mu_1}
(\frac{M_{s_0}}{M_P})^3 x_Q
\end{eqnarray}
\noindent
where we have used equation (\ref{uprime_lim}). This can be written in terms of, $N_Q$, the number of e-folds before the
end of inflation at which the Hubble radius crossed the de-Sitter
horizon. We have
\begin{eqnarray}
N_Q & = & -\frac{8 \pi}{M_P^2} \int_{Q}^{end}
\frac{U(s_0)}{U'(s_0)} ds_0 \nonumber \\
& \approx & \frac{32 \pi^3}{\mu_1^2}(\frac{M_{s_0}}{M_P})^2 (x^2_Q
-x^2_{end}) \nonumber\\
& \approx & \frac{32 \pi^3}{\mu_1^2}(\frac{M_{s_0}}{M_P})^2 x^2_Q
\label{efold}
\end{eqnarray}
\noindent
where we have used the approximation equation (\ref{uprime_lim}).
Hence
\begin{equation}
(\frac{\Delta T}{T})_{infl} \approx \sqrt{\frac{8\pi}{45}} \sqrt{ 8 \pi N_Q} (\frac{M_{s_0}}{M_P})^2
\end{equation}
The SU(2) phase transition produces cosmic strings which contribute
to the CMB anisotropy. The anisotropy due to strings can be
approximated as \cite{shellard}
\begin{equation}
(\frac{\Delta T}{T})_{string} \approx 9 G\mu
\end{equation}
\noindent
where $\mu$ is the string mass per unit length. We will approximate
\begin{equation}
\mu \approx M_{s_0}^2
\end{equation}
The combined temperature anisotropy is then \cite{jean1}
\begin{eqnarray}
(\frac{\Delta T}{T})_{tot} & \approx & \sqrt{ (\frac{\Delta T}{T})^2_{infl} +
(\frac{\Delta T}{T})^2_{string}} \nonumber\\
& \approx & \sqrt{81 + \frac{8 \pi}{45} \cdot 8\pi N_Q}
(\frac{M_{s_0}}{M_P})^2.
\end{eqnarray}
We will assume that $N_Q \approx 60$. From the CMB temperature
anisotropy value of $10^{-5}$ we find $M_{s_0} \approx
7 \cdot 10^{15}$GeV. We can estimate $\mu_1$ from equation (\ref{efold}).
If the Hubble radius crosses the de-Sitter horizon at $x_Q \approx 10$
we find $\mu_1 \approx 2 \cdot 10^{-2}$.
The spectral index is \cite{dvali1}
\begin{eqnarray}
n & = & 1 - 6 \epsilon + 2 \eta \nonumber\\
& \approx & 1 - \frac{1}{N_Q} = 0.98.
\end{eqnarray}
where we have used (\ref{ep_limit}) and the result $\mu_1 \ll 1$
\subsection{$U(1)$ Theory with Non-Zero Fayet-Illiopoulos Term}
Calculations of the CMB anisotropy for the $U(1)$ Fayet-Illiopoulos
case is identical to the $SU(2) \times U(1)$ case except the
string tension, $\mu$, is identified as \cite{lyth}
\begin{equation}
\mu \approx M_{\phi_-}^2 \approx 2 \pi \kappa
\end{equation}
\noindent
since the phase transition occurs at $\langle \phi_- \rangle =
\sqrt{\kappa} = M_{\phi_-}$.
The CMB anisotropy then constrains $\kappa \approx 8 \cdot 10^{30}$ (GeV)$^2$.
\section{Comments and Conclusions}
We have examined two SUSY models which lead to strings and found that
that the models inflate -- via hybrid inflation. The two models show
the different ways SUSY breaking can produce and end inflation -- by a
nonzero vev of a $F$ term, or a $D$ term. A $F$-term Abelian theory
would produce very similar results -- hence we have not included it in
the analysis. We believe hybrid inflation to be a somewhat generic
feature of SUSY unified theories. In fact, Jeannerot claims that
hybrid inflation is common to most unified theories which exclude
gravity with interactions of rank greater than or equal to five
\cite{jean2}. Our results together with \cite{jean1} generalise this
and show that supersymmetric theories with cosmic strings in general,
are hybrid inflation models.
The CMB anisotropy of our models was found to be of order
$(M_{s_0}/M_P)^2$. This is of the right magnitude, and allows GUT
scale physics to explain the CMB.
However, it is not clear how to extract the same inflation from a
higher energy theory such as supergravity or superstring theory. $F$
term models in supergravity tend to cause all scalar fields to be
renormalised, even the inflaton. This causes problems. It may
destabilise the flat directions and cause the slow-roll conditions to
be violated\cite{dvali2}. $D$ term supergravity models do not suffer
from this flaw. But any theory relying on $D$ term inflation must not
be semi-simple. It's gauge group must contain a $U(1)$ factor.
Further, to preserve the flat directions of global SUSY, all fields
charged by the Fayet-Illiopoulos $U(1)$ must be driven to zero. This
is not always possible. Also, in the context of superstring
theory, if the $D$-term dominates any $F$-term during inflation a
runaway dilaton problem may occur. \cite{lyth}
Nevertheless, we believe that SUSY hybrid inflation is sufficiently
attractive to warrant further investigation. We consider it
worthwhile to deduce whether hybrid SUSY inflation can easily fit
into a unified theory including gravity.
One caveat to our results is that the cosmic strings in the $D$-term
theory contain fermion zero modes that survive supersymmetry breaking.
Such strings could cause potential cosmological problems with the over
production of vortons\cite{vortons}. Also, in the usual cosmic
string scenario, the Abelian Higgs model is used as a prototype
theory. Recently this has been used in mixed string and inflation
scenarios \cite{joao,battye}. However, the evolution of cosmic
strings in realistic theories, such as those considered in this paper,
is an open question and deserves investigation.
\section*{Acknowledgements}
This work was supported in part by PPARC. M.M. is grateful to the
Cambridge Commonwealth Trust, Hughes Hall, and DAMTP for financial
support while this work was being completed.
|
\section{Introduction}
\label{intro}
Precision measurements play a significant role in contemporary
particle physics, spanning the energy scales starting with the
$Z$-boson pole at the high end, down to scales of the order of the $b$
and $c$ quark masses, and further down to hydrogen, positronium, and
muonium spectroscopy, and anomalous magnetic moments of leptons. The
role of the precision measurements reflects the successes
achieved by the Standard Model (SM) of electroweak and strong
interactions. It has become clear in recent years that finding any
signal of ``New Physics,'' beyond the SM expectations, is a very
difficult task.
Searching for hints about a more fundamental theory, which would
replace the SM, but lacking devices necessary to make the next step up
in energy, one resorts to a more precise analysis of the already
available data and tries to increase the accuracy of the experiments
at achievable energies. As a consequence, more and more accurate
theoretical predictions are required practically at every point of the
energy scale up to $\sim 100$ GeV.
Theoretical studies are often impeded by complications inherent to
bound states or QCD uncertainties. Consider for example the inclusive
decay width of the $B$-meson, $\Gamma(B \to X_c e\nu)$, used for the
$|V_{cb}|$ determination. To calculate the decay rate in a
model-independent way, one first has to clarify how to treat the
non-perturbative bound state effects. This non-trivial issue is
solved by invoking the Heavy Quark Expansion which permits a
calculation of the inclusive semileptonic decay width as a systematic
expansion~\cite{Bigi97} in powers of $\Lambda_{\rm QCD}/m$. The lowest
order of this expansion corresponds to the free quark decay. There
remains a technical difficulty of performing the standard perturbative
calculations. Until recently, only one-loop accuracy was achievable
for charged particle decays. Our approach to performing two-loop
calculations is one of the subjects of this talk.
A complicated situation occurs also at the very low energies, where
Coulombic bound states like hydrogen, muonium, or positronium are
studied. The measurements in such systems are very precise. The
theoretical predictions, however, are complicated by the essentially
non-perturbative nature of bound state calculations. Fortunately, it
turns out that the properties of the non-relativistic bound states,
such as energy intervals and lifetimes, can be described in terms of
series in powers and {\em logarithms} of the fine structure constant.
Here a proper organization of the calculations is a challenging
problem.
It is common in theoretical physics that problems can be solved if a
small parameter is available. In case of the ``standard''
perturbative calculations the small parameter is usually the coupling
constant. For the bound states, the smallness of a coupling
constant is insufficient, because the very existence of the system can
only be described in terms of an infinite number of Feynman diagrams.
The situation is simplified in case of non-relativistic bound states,
which encompass the light atoms and ``leptonia'' for which precision
measurements are being performed. There, the relative velocity of the
bound state components is small and can serve as an expansion
parameter.
For the bound states, expanding around a non-relativistic limit is the
only way known at present to arrive at precise results. But even in
case of the more conventional perturbative calculations, Feynman
integrals often are intractable and the only option to
evaluate them is to {\em invent} a small parameter. It is often
convenient to use some kinematic parameter, so that gauge
invariance and other symmetries of the problem are preserved.
Construction of non-conventional expansions of Feynman graphs has
recently helped to solve many outstanding problems in high precision
theoretical studies. In what follows we are going to describe some
examples of how such expansions can be constructed, and review our
recent calculations devoted to the semileptonic decay width of the $b$
quark and the positronium energy levels. We deliberately focus on the
technical aspects of those calculations, to provide additional
information and supplement our journal publications.
\section{Inclusive semileptonic decay width of the $b$ quark}
Let us start by discussing a calculation of the second order QCD
corrections, ${\cal O}(\alpha_s^2)$, to the semileptonic decay width
of the $b$ quark, $\Gamma(b \to c e \nu_e)$. At present a complete
calculation is not feasible. However, one can calculate corrections
at this order to the differential decay width ${\rm d}\Gamma/{\rm
d}q_{e \nu}^2$ for special values of the leptonic invariant mass,
$q_{e \nu}^2 \equiv q^2$. Such calculations were performed at three
points:
\begin{itemize}
\item
$q^2=(m_b-m_c)^2$ (zero recoil);~\cite{zerorecoil,zerorecoilA}
\item
$q^2=m_c^2$
(intermediate recoil);~\cite{b2cHalf}
\item
$q^2=0$ (maximal recoil).~\cite{Czarnecki:1997hc}
\end{itemize}
These kinematical configurations are depicted in
Fig.~\ref{fig:triangle}.
\begin{figure}
\hspace*{-19mm}
\begin{minipage}{16.cm}
\vspace*{3mm}
\[
\mbox{
\psfig{figure=triangle.ps,width=80mm,bbllx=25pt,bblly=104pt,%
bburx=582pt,bbury=630pt}
}
\]
\end{minipage}
\caption{Status of the two-loop QCD corrections to the decay $b\to
c+\mbox{leptons}$. The dashed line denotes the physical region for
the actual $c$ quark. Points where the full corrections are known are
circled. An analytical formula is known along the whole zero recoil
line.~\protect\cite{zerorecoilA,Franzkowski:1997vg} The other two
points are found from expansions around the base points of the two
arrows: at maximal recoil~\protect\cite{Czarnecki:1997hc} and at the
intersection with the diagonal.~\protect\cite{b2cHalf}}
\label{fig:triangle}
\end{figure}
Based on these 3 results, we performed a fit to interpolate the
corrections for arbitrary $q^2$. From that fit, a correction to the
inclusive decay rate, $\Gamma(b \to c e\nu_e)$, was estimated.
For kinematical configurations away from the zero recoil limit one
should consider three different types of corrections: pure virtual
corrections, single gluon radiation, and radiation of two gluons (at
zero recoil only the first type is present). Our approach is to
calculate these contributions separately and then sum them up in the
final result.\footnote{An alternative approach to such calculations is
based on the optical theorem; to get an $n$-loop result, one
calculates the imaginary part of a certain $(n+1)$-loop Feynman
integral. Such method was applied e.g.~to calculate QED corrections to
the muon decay rate.~\cite{vanRitbergen:1998yd}} Below we describe the
essential ingredients of those calculations.
\subsection{Virtual corrections}
\label{virtual}
As an example of the virtual correction calculation we consider the
intermediate recoil case, equating the invariant mass of the lepton
system, $\sqrt{q^2}$, to the $c$-quark mass. If the $c$ quark had
exactly half the $b$ quark mass, this configuration would coincide
with the zero-recoil limit (see Fig.~\ref{fig:triangle}), in which
case the virtual corrections are known
exactly.~\cite{zerorecoilA,Franzkowski:1997vg} For the actual $c$ and
$b$ masses these corrections can be calculated by expanding the
amplitude in power series in $(m_b-2m_c)/m_b$, that is around the
known zero recoil case. An important aspect of this expansion is that
it commutes with the virtual momenta integrations; it simply amounts
to Taylor expanding the integrand.
After the expansion, all integrals we need are of the
zero-recoil type. They can be divided up into the following categories
(they are written here in a general form, with $\omega$ denoting the
ratio of the final and the initial quark masses.
For our intermediate recoil calculation we need only $\omega=1/2$):
\begin{enumerate}
\item Planar with five propagators
\begin{eqnarray}
\lefteqn{P_5(a_1,a_2,a_3,a_4,a_5) = \pi^D }
\nonumber \\
&&\times\int
{ {\rm d}^D k_1 {\rm d}^D k_2
\over
k_1^{2a_1} (k_1-k_2)^{2a_2} k_2^{2a_3}
\left( k_1^2 + 2k_1p\right)^{a_4}
\left( k_2^2 + 2k_2p\omega \right)^{a_5}
}.
\label{eq:defPlanar}
\end{eqnarray}
\item Non-planar, first type, five propagators
\begin{eqnarray}
\lefteqn{N_P(a_1,a_2,a_3,a_4,a_5) = \pi^D
\int
{ {\rm d}^D k_1 {\rm d}^D k_2
\over
k_1^{2a_1} k_2^{2a_2} }}
\nonumber \\
&& \times \frac {1}{
\left( k_1^2 + 2k_1p\right)^{a_3}
\left( k_2^2 + 2k_2p\omega \right)^{a_4}
\left[ (k_1+ k_2)^2 + 2(k_1+k_2)p \right]^{a_5}
}
\label{eq:defNP}
\end{eqnarray}
\item Non-planar, second type, five propagators
\begin{eqnarray}
\lefteqn{N_P'(a_1,a_2,a_3,a_4,a_5) = \pi^D \int
{ {\rm d}^D k_1 {\rm d}^D k_2
\over
k_1^{2a_1} k_2^{2a_2} }
}
\nonumber \\
&&\times \frac {1}{
\left( k_1^2 + 2k_1p\right)^{a_3}
\left( k_2^2 + 2k_2p\omega \right)^{a_4}
\left[ (k_1+ k_2)^2 + 2\omega(k_1+k_2)p \right]^{a_5}
}
\label{eq:defNP1}
\end{eqnarray}
\item Non-planar with six propagators
\begin{eqnarray}
\lefteqn{N_6(a_1,a_2,a_3,a_4,a_5,a_6) = \pi^D
\int
{{\rm d}^D k_1 {\rm d}^D k_2
\over
k_1^{2a_1} k_2^{2a_2}
\left( k_1^2 + 2k_1p\right)^{a_3}}
}
\nonumber
\\
&&\hspace*{-10mm}
\times \frac {1}{
\left[ (k_1+ k_2)^2 + 2(k_1+k_2)p \right]^{a_4}
\left[ (k_1+ k_2)^2 +2\omega(k_1+k_2)p \right]^{a_5}
\left( k_2^2 +2\omega k_2p \right)^{a_6}
}
\nonumber \\
\label{eq:defN6}
\end{eqnarray}
\end{enumerate}
Taylor expansion of the amplitude may result in very many terms,
differing by powers of the propagators, $a_i$. It is crucial to find
a general algorithm of calculating them, so that an algebraic
manipulation program can be employed. It is convenient to use
recurrence relations derived using integration-by-parts,~\cite{che81} to
reduce an arbitrary integral to a small basis set of the so-called
master integrals.
Let us note at this point that solving recurrence relations is still
more of an ``art'' than ``science.'' It would be very desirable to
have a better understanding of the mathematical structure of the
systems of recurrence relations; a very promising approach to this
problem is being developed by P. Baikov.~\cite{Baikov:1996rk}
Here, we would like to demonstrate our approach to solving such
recurrence relations with the example of the planar integral
$P_5(\{a_i\})$. First, a system of relations is derived in the usual
way~\cite{che81} by equating integrals of total derivatives to zero.
One finds
\begin{eqnarray}
M_1 &=& D-2a_1-a_2-a_4+a_2{\bf 2}^+({\bf 3}^--{\bf 1}^-)-a_4{\bf 4}^+{\bf 1}^-
\nonumber\\[2mm]
M_2 &=&
D-a_1-2a_2-a_4+a_1{\bf 1}^+({\bf 3}^--{\bf 2}^-)
+a_4{\bf 4}^+\left({\omega-1\over\omega}{\bf 3}^--{\bf 2}^-
+{1\over \omega}{\bf 5}^-\right)
\nonumber\\[2mm]
M_3 &=& D-a_1-a_2-2a_4-a_1{\bf 1}^+{\bf 4}^-
+a_2{\bf 2}^+\left({\omega-1\over\omega} {\bf 3}^--{\bf 4}^-+{{\bf 5}^-\over\omega}\right)
+2a_4{\bf 4}^+
\nonumber\\[2mm]
M_4 &=& D-2a_2-a_3-a_5+a_3{\bf 3}^+({\bf 1}^--{\bf 2}^-)
+a_5{\bf 5}^+\left(
(1-\omega){\bf 1}^--{\bf 2}^-+\omega{\bf 4}^-\right)
\nonumber\\[2mm]
M_5 &=&
2D-2a_1-2a_2 -2a_3 -a_4-a_5 -a_4{\bf 4}^+{\bf 1}^--a_5{\bf 5}^+{\bf 3}^-
\nonumber\\[2mm]
M_6 &=&
D-a_2-a_3-2a_5+a_2{\bf 2}^+\left((1-\omega){\bf 1}^-+\omega{\bf 4}^--{\bf 5}^-\right)
\nonumber\\ && \qquad \qquad
-a_3{\bf 3}^+{\bf 5}^- +2\omega^2a_5{\bf 5}^+
\nonumber\\[2mm]
M_7 &=&
\left({3D\over 2}-\sum_{i=1}^5 a_i\right)({\bf 1}^--{\bf 4}^-)
+2a_4{\bf 4}^+{\bf 1}^-+a_5{\bf 5}^+ \omega({\bf 1}^--{\bf 2}^-+{\bf 3}^-)
\nonumber\\[2mm]
M_8 &=&
{1\over \omega}
\left({3D\over 2}-\sum_{i=1}^5 a_i\right)({\bf 3}^--{\bf 5}^-)
+a_4{\bf 4}^+({\bf 1}^--{\bf 2}^-+{\bf 3}^-) +2\omega a_5{\bf 5}^+ {\bf 3}^-
\label{eq:system}
\end{eqnarray}
The recursive algorithm proceeds in three steps:
\begin{enumerate}
\item
the exponents of the massive propagators ($a_{4,5}$) are reduced to 1;
\item
the exponent of the massless propagator containing both integration
impulses ($a_2$) is reduced to 1;
\item
the exponents $a_{1,3}$ are reduced.
\end{enumerate}
The first step is simple. We assume that $a_4$ and $a_5$ are
positive; otherwise the integral can be done by a sequence of one-loop
integrations. We use the relation $M_6$.
The term $2\omega^2 a_5{\bf 5}^+$ involves an integral with $a_5$ higher
than any other term in this relation. Therefore we can use $M_6$
to express $P_5(\ldots,a_5+1)$ by $P_5(\ldots,a_5)$.
An analogous manipulation is done using $M_3$ to reduce $a_4$.
Eventually we obtain $a_4,a_5 \le 1$.
The next step is accomplished as follows.
We first would like to eliminate all operators that contain ${\bf 4}^+$ and
${\bf 5}^+$ from the system (\ref{eq:system}). There are eight such
operators ${\bf 4}^+$, ${\bf 4}^+{\bf 1}^-$, ${\bf 4}^+{\bf 2}^-$, ${\bf 4}^+{\bf 3}^-$, and ${\bf 5}^+$,
${\bf 5}^+{\bf 1}^-$, ${\bf 5}^+{\bf 2}^-$, ${\bf 5}^+{\bf 3}^-$. An important point is that
this number is equal to the number of the recurrence relations and we
can therefore solve the recurrence relations for these operators.
We denote the resulting 8 equations by $T_{4,41,42,43,5,51,52,53}$.
The notations are such that for example $T_{41}$ means an equation
which expresses the operator ${\bf 4}^+{\bf 1}^-$ through other operators. These
equations allow us to obtain new recurrence relations using
$a_4( T_{41} - {\bf 1}^- T_4) = 0$ and similar equations. In
particular, we use a relation $a_4( T_{43} - {\bf 3}^- T_4)=0$ to
lower the value of $a_2$ to $1$.
After this has been achieved, we eliminate all operators which contain
${\bf 2}^+$. At this stage, the problem is essentially two parametric
(i.e.~we have to move in the $a_1-a_3$ plane). Finally, we are able
to reduce any $P_5(\{a_i\})$ integral to four integrals
$P_5(1,1,1,1,1)$, $P_5(0,1,1,1,1)$, $P_5(-1,1,0,1,1)$, and
$P_5(0,1,0,1,1)$ and a number of trivial integrals. The four master
integrals are
\begin{eqnarray}
P_5(1,1,1,1,1)&=&-\frac {\pi^2}{3\epsilon} -\frac {2 \pi^2}{3}
- \frac{5\zeta (3)}{2} - \pi^2\ln 2
\nonumber \\
P_5(0,1,1,1,1) &=& \frac {1}{2\epsilon^2} + \frac {1}{\epsilon}\left( \frac {5}{2}
+ 2\ln2 \right ) + \frac {19}{2} + 10\ln 2
- \frac {5}{12}\pi^2
\nonumber \\
P_5(-1,1,0,1,1) &=&
\frac {19}{16\epsilon^2}
+ \frac {1}{\epsilon} \left ( \frac {311}{96} + \frac {3\ln 2}{4}
\right )
+ \frac {1225}{192} + \frac {79}{24}\ln 2 - \frac {\ln^2 2 }{2}
+ \frac {15 \pi^2}{32}
\nonumber \\
P_5(0,1,0,1,1) &=&
- \frac {5}{8\epsilon^2}
- \frac {1}{\epsilon} \left ( \frac {29}{16} + \frac {\ln 2}{2} \right)
- \frac {107}{32} - \frac {13}{4}\ln 2 + \ln^2 2
- \frac {11\pi^2 }{48}.
\nonumber
\end{eqnarray}
The above approach to solving recurrence
relations is quite general. One sees that such solution cannot be
constructed in a
completely universal way. However, our approach allows to go through
a number of steps, constructing a solution without much effort and it
has proved useful in various cases. In particular, it works for the
``hard'' threshold integrals, necessary for the calculations described
in Section \ref{sec:hard}.
\subsection {Emission of two gluons in semileptonic $b$ decays}
For simplicity, we will discuss here the case of the maximal recoil,
$q^2=0$. The intermediate recoil case is treated in an analogous way.
A more detailed discussion of this issue is given in
\cite{maxtech}. Here the most essential features are summarized.
The basic observation in this part of the calculation is that the
energy losses caused by radiation off a non-relativistic particle are
small. This is evident in the Coulomb gauge, since the interaction of
magnetic photons with charged particles is proportional to the
velocity of the latter. Therefore, the physical properties of the real
radiation provide a hint how an expansion should be organized.
Considering a decay $b \to c +e + \nu_e +g_1 + g_2$ in the case
$q^2\equiv (e+\nu_e)^2 =0$, we intend to calculate it expanding in
$\delta = (m_b-m_c)/m_b$. For brevity we will refer to the $e,\nu$
system as to the $W$ boson. Taking into account phase space
constraints, it is easy to establish that the spatial momentum of the
$c$ quark, as well as the four-momenta of the $W$, $g_1$ and $g_2$ are
of the order $\delta$. Then, a natural expansion of the propagators
emerges: all propagators of the virtual particles can be expanded
around the {\em static limit}, which essentially simplifies the
calculation. Clearly, such an expansion will generate a large number
of terms, but this can be handled using symbolic manipulation
programs.
The last non-trivial question is the integration of the obtained
expression over the four particle phase space. Given that the
remaining propagators are static and the $W$ boson does not
participate in the interaction, it is clear that the integration over
two-particle phase (sub)space ($W,c$) can be easily performed. The
remaining part of the calculation requires an integration over the
three particle phase space -- two gluons and the ``particle'' composed
of $W$ and $c$. A systematic expansion of this object is accomplished
by an appropriate choice of variables, described in detail in
\cite{maxtech}.
\subsection{One gluon radiation and one virtual correction}
Single gluon emission in the presence of a virtual loop is, in
principle, possible to compute exactly. However, it is very inconvenient,
e.g.~because of necessity of integrating four-point functions over a
three particle phase space.
We have adopted a different approach, based again on an expansion of
such amplitudes around the static limit ($m_b=m_c$). In contrast to
the virtual corrections and double gluon emission, in the present case
the expansion is non-trivial. Let us consider the following diagram:
first a virtual gluon is exchanged between $c$ and $b$ quarks, and
then the $c$ quark emits a real gluon. An essential point is that
before the real gluon emission the $c$ quark is {\em off mass shell},
which provides an infra-red regularization for the virtual loop.
Expanding the one-loop amplitude naively, i.e.~putting the
``intermediate'' $c$-quark line loop on shell, we change the infrared
properties of the diagram and divergences will result in singularities
in $1/(D-4)$ (we use dimensional regularization). To avoid these
spurious singularities the expansion should be ``corrected''; in this
case the so-called eikonal expansion~\cite{CzarSmir96,Smirnov96} is
used. In the context of the two loop corrections to fermion decays
this expansion is described in detail in Ref.~\cite{maxtech}.
\subsection{ General remarks on semileptonic heavy fermion decays}
Let us summarize the technical experience we have gained by studying
semileptonic $b$ and $t$ decays to second order in the strong coupling
constant.
First, we note that the problem was made tractable by inventing a ``small''
parameter. For the top quark decays~\cite{Czarnecki:1998qc} the
result is obtained if we take the ``small'' parameter close
to unity. This clearly pushes the applicability of the method to its
limits. However, even in this extreme case reasonable numerical
estimates can be obtained.
Second, because we consider processes with the on--shell external
particles, one has to be careful in analyzing contributing regions of
integration momenta. In particular, one can encounter non-analytical
dependence on the expansion parameter. In many cases the relevant
momentum regions can be found by considering the corresponding
effective theories.
Third, it is possible to construct an expansion of the amplitudes with
real radiation of massless particles starting from the
non-relativistic limit. It is likely that this construction can be
applied beyond the semileptonic decay width, for example in a
perturbative calculation of the cross section of such process as
$e^+e^- \to \mu^+ \mu^- \gamma \gamma$ close to the two muon threshold,
or even for higher energies if one takes into account sufficiently
many terms of such an expansion.
\section{Bound States in QED}
Consistent description of bound states is one of the most complicated
problems in Quantum Field Theory. The problem is simplified if the
bound state components are non-relativistic. This is the case, for
example, in QCD in $\Upsilon$ or $\psi$ meson families, or in QED in
light atoms and in leptonic bound states like positronium or muonium.
The difficulty in the study of bound states is their essentially
non--per\-tur\-bative nature, as we discussed in Section \ref{intro}.
In the non-relativistic case the situation is simplified: to first
approximation one can describe the bound state using a Schr\"odinger
wave function. This description can subsequently be refined by means
of perturbation theory, using the relative velocity $v$ of the
components as an expansion parameter.
This kind of expansions can be constructed both for QED and QCD. In
what follows, we limit ourselves to the QED where a low-energy
effective theory approach is well established, and is known as
Non-Relativistic Quantum Electrodynamics (NRQED).~\cite{Caswell:1986ui}
\subsection{NRQED}
Very roughly speaking, the idea underlying NRQED consists in
separating the soft scale $\sim m\alpha$, which characterizes the
bound state, from the hard one $\sim m$, and ``integrating out'' the
latter. The interactions at the soft scale can be described by
Quantum Mechanics, i.e.~employing a non-relativistic approximation.
The physics at hard scales is governed by the relativistic QFT.
However, when we are interested in processes which occur at small
momenta and energies, the contribution of hard scales can be accounted
for by introducing effective local operators and computing their
coefficients in the effective Hamiltonian.
In the lowest order of this construction we describe the system by a
bound state wave function. For positronium it is found from
the Schr\"odinger equation with the ``leading order'' Hamiltonian ($m$
is the electron mass, $r$ is the distance between the electron and the
positron):
\begin{equation}
H_{\rm LO} = \frac {{\bf p} ^2}{m}-\frac {\alpha}{r}.
\label{eq:loh}
\end{equation}
In a 3-dimensional space the ground state wave function (in momentum
representation) is
\begin{equation}
\phi(p) = \sqrt{\pi\alpha m\over 2}{2m^2\alpha^2\over ({\bf p}^2-mE)^2},
\qquad
E=-{m\alpha^2\over 4}.
\label{wf}
\end{equation}
The difference between the full Hamiltonian and Eq.~(\ref{eq:loh}) can
be treated as a perturbation. The difficulties of calculating in
NRQED are caused by divergences in the matrix elements of the
subleading operators. In the original formulation~\cite{Caswell:1986ui}
the ultraviolet divergences were dealt with by subtracting
counterterms defined with help of diagrams evaluated on-shell and at
threshold. Those diagrams, in turn, contain infrared divergences,
which are regularized by introducing a small mass $\lambda$ for the
photon.
Since that pioneering work it has become clear that working with a
massive photon leads to technical difficulties, particularly in
evaluating the so-called hard corrections, arising at scales $\sim m$.
Indeed, the hard corrections remain unknown (at two-loops) for such
processes as para or orthopositronium decays, while the soft scale
effects have already been evaluated. The reason for this is that the
two-loop diagrams are difficult to evaluate when more than one mass
scale is present. Although the only relevant scale in hard
corrections for positronium physics is the electron mass,
regularization with a photon mass introduces a second scale $\lambda$.
The hard corrections can, however, be computed in many configurations,
if instead of introducing $\lambda$ one uses dimensional
regularization.
It has been shown~\cite{Pineda:1997bj} that dimensional regularization
permits an exact separation of effects arising at various
characteristic energy scales. Using that method, the complete energy
spectrum of Ps has been reproduced~\cite{Pineda:1998kn} to order
$m\alpha^5$. Further, we have computed analytically $m\alpha^6$
corrections to the hyperfine splitting (HFS) of the Ps ground
state.~\cite{Czarnecki:1998zv} More recently,~\cite{levels} we
generalized that result to all $S$ states and computed also their spin
independent shift at ${\cal O}(m\alpha^6)$. Here we would like to
discuss some features of the NRQED in dimensional regularization,
which we found useful in those calculations.
\subsection{Quantum Mechanics in $d$ dimensions}
The first point we would like to discuss is a construction
of the lowest order approximation. We
consider the Schr\"odinger
equation in a $d$-dimensional momentum space,
\begin{equation}
\phi(p) = \fr {4\pi \alpha m}{{\bf p}^2-mE}
\int \fr {{\rm d}^{\rm d} {\bf k}}{(2\pi)^d}
\fr {\phi(k)}{({\bf p}-{\bf k})^2}.
\label{ddS}
\end{equation}
The solution of this equation for arbitrary $d$ is not known.
This complicates the calculations which involve the wave function.
However, since in the configuration space all divergences
are located at $r=0$, it is possible to extract divergent
contributions in the form of $\psi^2(0)/\epsilon$. Thus, the unknown
overall factor
$\psi^2(0)$ will be present in the intermediate results. In the final
result the singularities $1/\epsilon$ cancel and we can replace $\psi(0)$
by its 3-dimensional value.
In rare cases we need, in addition to the value of the configuration
space wave function at the origin, also the behavior of its first
derivative. This, however, is related to $\psi(0)$ by the
Schr\"odinger equation. To demonstrate this, we consider
the Schr\"odinger equation in the configuration space. Since for $S$
states $\psi(r) $ has no angular dependence, we have
\begin{eqnarray}
\left( -{\nabla^2\over m} + V(r)\right) \psi(r)
= -{1\over m}\left(
\partial_r^2 + {d-1\over r} \partial_r - m V(r)\right) \psi(r)
= 0,
\label{eq:ddS}
\end{eqnarray}
where $V(r)$ is the Coulomb potential in $d$ dimensions,
\begin{equation}
V(r) = - {\alpha \Gamma\left({d\over 2} -1\right) \over
\pi^{{d\over 2}-1} r^{d-2}}.
\end{equation}
We assume that the wave function near the origin behaves like
$
\psi(r) \simeq \psi(0) ( 1-cr^\beta),
$
and find $c$ and $\beta$ by substituting this form into
Eq.~(\ref{eq:ddS}):
\begin{equation}
\psi(r\to 0) \simeq \psi(0) \left( 1
- {m\alpha\Gamma\left({1\over 2} - \epsilon\right)
\over 2\pi^{{1\over 2}-\epsilon}(1+2\epsilon)}r^{1+2\epsilon}
\right).
\end{equation}
We now would like to demonstrate how divergent integrals are treated
in practice, using two examples from our hyperfine splitting
calculations. First, we consider a logarithmic divergence
arising from tree-level Coulomb photon exchange,~\cite{levels}
\begin{equation}
\Delta_C E_{\rm hfs}
= - \fr{\pi\alpha}{dm^4}
\left\langle \fr{ ({\bf p}'{\bf q})({\bf q}{\bf p}) }{ {\bf q}^2 } \right\rangle.
\label{int}
\end{equation}
In Eq.~(\ref {int}) the matrix element is to be calculated over the ground
state wave function in $d$ dimensions:
$$
\left\langle f({\bf p},{\bf p}') \right\rangle \equiv
\int {d^Dk_2^d p \over (2\pi)^d}{d^Dk_2^d p' \over (2\pi)^d}
\phi(p)\phi(p') f({\bf p},{\bf p}').
$$
Although the integrand does not look complicated, the
difficulty is that the exact form of the wave function $\psi(r)$ in
$d$ dimensions is not known. We use the fact that it
satisfies the $d$-dimensional Schr\"odinger
equation, Eq.~(\ref{ddS}).
Using that equation we rewrite the integral in Eq.~(\ref {int}) as
\begin{equation}
\left\langle \fr{ ({\bf p}'{\bf q})({\bf q}{\bf p}) }{ {\bf q}^2 } \right\rangle _{{\bf p},{\bf p}'} =
\left\langle \fr{(4\pi\alpha m)^2({\bf p}'{\bf q})({\bf q}{\bf p}) }
{({\bf p}^2-mE)({\bf p}-{\bf k})^2 {\bf q}^2 ({\bf p}'^2-mE)({\bf p}'-{\bf k}')^2}
\right\rangle _{{\bf k},{\bf k}'},
\label {int1}
\end{equation}
where the integration over ${\bf p}$, ${\bf p}'$, as well as ${\bf k}$, ${\bf k}'$, in the
last expression is
understood. The integral over ${\bf p}$ and ${\bf p}'$ receives a divergent
contribution only from the region where ${\bf p}$ and ${\bf p}'$ become
simultaneously infinite. Therefore, a single subtraction is sufficient
to make this integral finite. It is convenient to subtract from
(\ref{int1}) the following expression:
\begin{equation}
\left\langle \fr{(4\pi \alpha m)^2 ({\bf p}'{\bf q})({\bf q}{\bf p}) }{({\bf p}^2-mE)^2 {\bf q}^2
({\bf p}'^2-mE)^2} \right\rangle _{{\bf k},{\bf k}'}.
\label {int2}
\end{equation}
After the subtraction is done, two nice features emerge. In Eq.~(\ref
{int2}) the integration over ${\bf k},{\bf k}'$ factorizes and leads to
$\psi^2(0)$ times a two-loop integral, which can be easily calculated
for arbitrary $d$. On the other hand, the difference between the last
integral in Eq.~(\ref {int1}) and the integral in Eq.~(\ref {int2}) is
finite and can be calculated for $d=3$ using the explicit form of the
wave function, Eq.~(\ref{wf}).
We note that the counterterm (\ref {int2}) is
constructed in such a way that the difference mentioned above vanishes
for the ground state. This can be easily seen by integrating over
${\bf k},{\bf k}'$ in Eq.~(\ref {int2}) and noticing that the
${\bf p},{\bf p}'$-dependent terms in the denominator of Eq.~(\ref {int2})
coincide (up to a normalization factor) with the three-dimensional
ground state wave functions in the momentum representation.
Finally, for the Coulomb exchange contribution one finds
\begin{equation}
\Delta_C E_{\rm hfs} = \fr{ \pi\alpha^3 }{ 24 m^2 } \psi^2(0)
\left ( \fr{ 1 }{\epsilon } - 4 \ln(m\alpha) - \fr{1}{3} \right ),
\label{DeltaC}
\end{equation}
Our second example is a linearly divergent integral arising from the
tree-level magnetic photon exchange,~\cite{levels}
\begin{eqnarray}
\left\langle {\bf p}^2 \right\rangle &=& \psi(0)
\int \fr{d^Dk_2^d p}{(2\pi)^d} {\bf p}^2 \phi(p)
\nonumber \\
&=& m \psi(0)
\int \fr{d^Dk_2^d p}{(2\pi)^d} \left ( E \phi(p)
+
\int \fr{d^Dk_2^d k}{(2\pi)^d} \fr{ 4\pi\alpha }{ ({\bf p}-{\bf k})^2 }
\phi(k) \right ).
\label {tadpole}
\end{eqnarray}
Shifting the integration variable ${\bf p} \to {\bf p}+{\bf k}$ we find that the
${\bf p}$-integral in the last term is scale-less. In dimensional
regularization such integrals vanish. The first term in Eq.~(\ref
{tadpole}) is finite in three dimensions. We obtain
\begin{equation}
\left\langle {\bf p}^2 \right\rangle = mE \psi^2(0).
\end{equation}
\subsection{Hard scale contributions}
\label{sec:hard}
Perhaps the main advantage of using dimensional regularization in
bound state calculations is the possibility of computing
the hard corrections analytically.
The contributions of hard scales arise from virtual momenta of
order of the electron mass. They can be calculated by considering
the on--shell $e^+e^-$ scattering amplitude exactly at the
threshold, i.e.~for zero relative velocity of the incoming electron
and positron. Since the amplitude is independent
of the incoming spatial momenta,
it gives rise to four-fermion operators in the low-scale Lagrangian or,
equivalently, to the $\delta(\mbox{\boldmath$r$})$ terms in the
effective quantum mechanical Hamiltonian.
Technically, this calculation is similar to the derivation of
the matching coefficient of the vector quark-antiquark current in QCD
and its NRQCD counterpart.~\cite{threshold,BSS}
An arbitrary Feynman integral which contributes to the hard scale part of
the calculation can be written as
\begin{equation}
I(a_1,...a_9) = \int \fr {d^Dk_2^D k_1}{(2\pi)^D} \fr {d^Dk_2^D k_2}{(2\pi)^D}
{1\over S_1^{a_1}S_2^{a_2}S_3^{a_3}S_4^{a_4}S_5^{a_5}S_6^{a_6}S_7^{a_7}
S_8^{a_8} S_9^{a_9}},
\end{equation}
where
\begin{eqnarray}
&&
S_1 = {k_1^2},\qquad
S_2={k_2^2},\qquad
S_3={(k_1-k_2)^2}, \qquad
S_4 = {k_1^2+2pk_1},
\nonumber \\
&&
S_5= {k_2^2+2pk_2},\qquad
S_6 = {k_1^2-2pk_1},\qquad
S_7 = {k_2^2-2pk_2},
\nonumber \\
&&
S_8 = {(k_1-k_2)^2+2p(k_1-k_2)},\qquad
S_9 = {(k_1-k_2)^2-2p(k_1-k_2)},
\end{eqnarray}
and $a_1,\ldots,a_9$ are integers.
In practice we find diagrams with only at most 6 different
propagators, so that at least 3 exponents $a_i$ are zero.
A method for computing these integrals is analogous to the 2-loop
virtual corrections to quark decays. We have described our algorithm
in Section \ref{virtual}.
\section{Conclusions}
In this talk we have presented technical aspects of our recent
calculations of two-loop corrections to semileptonic quark decays and
positronium properties. The methods developed in the context of quark
decays, in particular the algorithm for solving systems of recurrence
relations, have found applications in positronium
physics. We hope that further progress can be made in such atomic
physics calculations using the tools developed for high energy
problems. In particular, applying dimensional regularization should
facilitate so far intractable calculations, such as two-loop
corrections to positronium decays.
|
\vec\partial{\stackrel{\rightarrow}{\partial}}
\def\vec\partial{\vec\partial}
\def\check\delta\!\!\!/{\check\delta\!\!\!/}
\def\begin{equation}{\begin{equation}}
\def\end{equation}{\end{equation}}
\def\begin{eqnarray}{\begin{eqnarray}}
\def\end{eqnarray}{\end{eqnarray}}
\def\begin{eqnarray*}{\begin{eqnarray*}}
\def\end{eqnarray*}{\end{eqnarray*}}
\renewcommand{\theequation}{\thesection.\arabic{equation}}
\begin{document}
\hfill {\bf UFRJ-IF-NCG-1999/2}
\vskip 1cm
\begin{center} {\Large \bf Connes-Lott model building on the two-sphere}
\vskip 0,5cm {\bf \large J.A.Mignaco${}^{(*)}$
\footnote {partially supported by CNPq,Brasil}, C.Sigaud${}^{(*)}$, A.R.da Silva${}^{(**)}$ and
F.J.Vanhecke${}^{(*)}$\\ {\small E-mail: <EMAIL>, <EMAIL>, <EMAIL>
and
<EMAIL> }}
\vskip 0.5cm
${}^{(*)}${\it Instituto de F\'\i sica}, ${}^{(**)}${\it Instituto de Matem\'atica,}\\ {\it UFRJ,
Ilha do Fund\~ao, Rio de Janeiro, Brasil.}\\
\end{center}
\begin{abstract}
In this work we examine generalized Connes-Lott models, with ${\bf C}\oplus{\bf C}$ as finite algebra, over the two-sphere.
The Hilbert space of the continuum spectral triple is taken as the space of sections of a twisted
spinor bundle, allowing for nontrivial topological structure (magnetic monopoles).
The finitely generated projective module over the full algebra is also taken as topologically
non-trivial, which is possible over $S^2$. We also construct a real spectral triple enlarging this
Hilbert space to include "particle" and "anti-particle" fields.
\end{abstract}
\vskip 4cm
\noindent PACS numbers : 11.15.-q, 02.40.-k ; \\ Keywords : Noncommutative geometry, Connes-Lott
model, Real spectral triples
\newpage
\section{Introduction}
\setcounter{equation}{0} In previous work \cite{Migetal1}, we studied the Connes-Lott program with
the complex algebra ${\cal A}_1={\cal C}(S^2;{\bf C})$ of continuous complex-valued functions on the sphere.
The Hilbert space ${\cal H}_I$, on which this algebra was represented, consisted of one of the minimal
left ideals ${\cal I}_{\pm}$ of the algebra of sections of the Clifford bundle over $S^2$ with a
standard scalar product. On this Hilbert space the Dirac operator was taken as
${\cal D}_I={\bf i}({\bf d}-\delta)$ restricted to each of the ideals ${\cal I}_\pm$. The projective modules
${\cal M}$ over ${\cal A}_1$ were constructed using the Bott projector
${\bf P}={1\over 2}({\bf 1}+\vec n\cdot\vec\sigma)$, acting on the free module
${\cal A}_1\oplus{\cal A}_1$. These modules are classified by the homotopy classes of the mappings $\vec
n:S^2\rightarrow S^2$ i.e. by $\pi_2(S^2)={\bf Z}$. In Dirac's interpretation, each integer corresponds
to a magnetic monopole at the center of the sphere with magnetic charge $g$ quantised by
\(eg/4\pi=(n/2)\hbar\).
In the present paper we extend the above analysis of topologically non trivial aspects in
noncommutative geometry, to the product algebra ${\cal A}={\cal A}_1\otimes{\cal A}_2$, where ${\cal A}_2={\bf C}\oplus{\bf C}$. It is
clear that ${\cal A}\simeq{\cal C}(S^2\times\{a,b\};{\bf C})$, where $\{a,b\}$ denotes a two-point space, as in
the original Connes-Lott paper \cite{C-L}. \\ In section {\bf \ref{Pensov}} we construct the
Hilbert space on which ${\cal A}_1$ is represented. This Hilbert space ${\cal H}_{(s)}$ generalizes ${\cal H}_I$
above, and is made of sections of what we call a Pensov spinor bundle of (integer or semi-integer)
weight $s$, using a taxonomy introduced by Staruszkiewicz \cite{Star}. A generalized Dirac operator
${\cal D}_{(s)}$ acting on these Pensov spinors is defined and, for $s=\pm1/2$ we recover the K\"{a}hler
spinors of \cite{Migetal1}, while for $s=0$ the usual Dirac spinors are obtained. These spinors may
actually be identified with sections of twisted spinor bundles or, from a more physical viewpoint, as
usual Dirac spinors interacting with a magnetic monopole of charge $g$ given by \(eg/4\pi\hbar=s\).\\
The projective modules over ${\cal A}$, following Connes-Lott, are constructed in
section {\bf\ref{Module}} as ${\cal M}={\bf P}({\cal A}\oplus{\cal A})$ where ${\bf P}=\Bigl({\bf P}_a={\bf 1},{\bf
P}_b={1\over 2}({\bf 1}+\vec n_b\cdot\vec\sigma)\Bigr)$. In Connes' work \cite{Con1}, the smooth
manifold is four-dimensional so that, taking the four-sphere $S^4$ as an example, we get mappings $\vec n :
S^4\rightarrow S^2$ classified by $\pi_4(S^2)={\bf Z}_2$. However, the local unitary
transformations acting as ${\bf P}_b\rightarrow U^\dagger{\bf P}_b U$ are also classified by homotopy classes
$\pi_4(U(2))=\pi_4(SU(2))=\pi_4(S^3)={\bf Z}_2$. It follows that\footnote{We are indebted to
prof.Balachandran of Syracuse University for discussions on this point.} all Bott projectors define
modules isomorphic to the module obtained from ${\bf P}_b={1\over 2}({\bf 1}+\sigma_3)$, considered
by Connes. For the two-sphere $S^2$ this does not happen since $\pi_2(S^2)={\bf Z}$ and
$\pi_2(U(2))=\pi_2(SU(2))=\pi_2(S^3)=\{{\bf 1}\}$. In section {\bf\ref{Triple}} we construct the full
spectral triple obtained as the product of the Dirac-Pensov triple for the algebra ${\cal A}_1$ with a
discrete spectral triple for the algebra ${\cal A}_2$. Following again Connes' prescription \`{a} la
lettre, we take the discrete Hilbert space as ${\cal H}_{dis}={\bf C}^{N_a}\oplus{\bf C}^{N_b}$ with chirality
$\chi_{dis}$ given as +1 on the a-sector and -1 on the b-sector. The discrete Dirac operator
${\cal D}_{dis}$ is then the most general hermitian matrix, odd with respect to the grading defined by
$\chi_{dis}$. Eliminating the "junk" in the induced representation of the universal differential
enveloppe $\Omega^{\bullet}({\cal A})$, yields bounded operators $\Omega^{\bullet}_D({\cal A})$ in
${\cal H}={\cal H}_{(s)}\otimes{\cal H}_{dis}$. The standard use of the Dixmier trace and of Connes'trace theorem
allows then to define a scalar product of operators in $\Omega^{\bullet}_D({\cal A})$. This scalar
product is used in section {\bf\ref{Yang-Mills-Higgs}} to construct the Yang-Mills-Higgs action.
The main new features in this action, as compared with Connes' result, are the
appearance of an additional monopole potential of strength \(eg/4\pi=(n/2)\hbar\), where $n$ is the
integer characterizing the homotopy class of ${\bf P}_b$, and the fact that the Higgs doublet is not
globally defined on $S^2$ but transforms as a Pensov field of weight $\pm n/2$. The particle sector
is examined in section {\bf\ref{Matter}} and a covariant Dirac operator ${\cal D}_\nabla$ acting on
${\cal H}_p={\cal M}\otimes_{\cal A}{\cal H}$ is defined. Here, the novelty is that, whilst the "a-doublet" continues
as a doublet of Pensov spinors of weight $s$, the "b-singlet" metamorphoses in a Pensov spinor of
weight $s+n/2$. If one should insist on a comparison with thestandard electroweak model on $S^2$,
this would mean that right-handed electrons see a different magnetic monopole than the left-handed
and this is not really welcome. In section {\bf\ref{Real Triples}} we introduce a real Dirac-Pensov
spectral triple by doubling the Hilbert space as ${\cal H}_{1}={\cal H}_{(s)}\oplus{\cal H}_{(-s)}$. It is seen that,
with the same ${\cal H}_{dis}$ as before, it is not possible to define a real structure. However, a more
general discrete Hilbert space ${\cal H}_{2}={\bf C}^{N_{aa}}\oplus{\bf C}^{N_{ab}}\oplus{\bf C}^{N_{ba}}\oplus{\bf C}^{N_{bb}}$,
as considered in\cite{Kraj,Pasch}, allows for the construction of a real structure on
${\cal H}_{new}={\cal H}_{1}\otimes{\cal H}_{2}$. The covariant Dirac operator on
${\cal M}\otimes_{\cal A}{\cal H}_{new}\otimes_{{\cal A}}{\cal M}^*$ can also be defined and it is furthermore seen that,
with the use of such a non trivial projective module, the abelian gauge fields are not slain,
as they are when ${\cal M}={\cal A}$\cite{Var}.
Clearly this model building led us far from a toy electroweak model. The main purpose however
is not to reproduce such a model on the two-sphere, but rather to examine some of the
topologically nontrivial structures in model building with the simplest manifold allowing for such
possibilities.
\newpage
\section{The Hilbert space of Pensov spinors on $S^2$}
\label{Pensov}\setcounter{equation}{0}
The standard atlas of the two-sphere $S^2=\{(x,y,z)\in {\bf R}^3 \mid x^2+y^2+z^2=1\}$ consists of two
charts, the boreal, $H_B=\{(x,y,z)\in S^2\mid -1<z\leq +1\}$, and austral chart,
$H_A=\{(x,y,z)\in S^2\mid -1\leq z < +1\}$, with coordinates :
\begin{eqnarray*}
\zeta_B&=&\xi^1_B+{\bf i} \xi^2_B=+ {x+{\bf i} y \over 1+z}\;\hbox{ in $H_B$}\;,\\
\zeta_A&=&\xi^1_A+{\bf i} \xi^2_A=- {x-{\bf i} y \over 1-z}\;\hbox{ in $H_A$}\;.
\end{eqnarray*}
In the overlap $H_B\cap H_A$, they are related by $\zeta_A\,\zeta_B=-1$ and the usual spherical
coordinates $(\theta,\varphi)$, given by $\zeta_B=-1/\zeta_A=\tan \theta/2\;\exp {\bf i}\varphi$, are
nonsingular. In each chart, dual coordinate bases of the complexified tangent and cotangent spaces
are :
\begin{eqnarray*}
&&\biggl\{\vec\partial={\partial\over\partial\zeta}= {1\over 2}\biggl({\partial\over\partial\xi^1}
-{\bf i}{\partial\over\partial\xi^2}\biggr) ,\vec\partial^{\;*}={\partial\over\partial\zeta^*}=
{1\over 2}\biggl({\partial\over\partial\xi^1}+{\bf i} {\partial\over\partial\xi^2}\biggr)\biggr\}\\
&&
\biggl\{d\zeta=d\xi^1+{\bf i} d\xi^2,d\zeta^*=d\xi^1-{\bf i} d\xi^2\biggr\}\;.\end{eqnarray*}
In $H_B\cap H_A$ they are related by
\begin{eqnarray*}
\left(
\begin{array}{cc}
\vec\partial_A & \vec\partial_A^{\;*} \end{array}
\right)
& = & \left(\begin{array}{cc}
\vec\partial_B & \vec\partial_B^{\;*}\end{array}\right)\;
\left(
\begin{array}{cc}
{\zeta_B}^2 & 0 \\
0 & {\zeta_B^*}^2\end{array}\right)\;,\\
\left(\begin{array}{c}
d\zeta_B \\ d\zeta_B^*\end{array}\right)
& = &
\left(\begin{array}{cc}
{\zeta_B}^2 & 0 \\
0 & {\zeta_B^*}^2\end{array}\right)
\left(\begin{array}{c}
d\zeta_A \\ d\zeta_A^*\end{array}\right)\;.\end{eqnarray*}
The euclidean metric in ${\bf R}^3$ induces a metric on the sphere :
\[{\bf g}={4\over q^2}\delta_{ij}\,d\xi^i\otimes d\xi^j=
{2\over q^2}\biggl(d\zeta^*\otimes d\zeta + d\zeta\otimes d\zeta^*\biggr)\;,
\]
where $q= 1+\vert\zeta\vert^2$.
Real and complex Zweibein fields are given by:
\begin{equation}\label{Hil1}
\biggl\{\theta^i={2\over q}d\xi^i\;;i=1,2\biggr\}\;\hbox{and}\;
\biggl\{\theta={2\over q}d\zeta,\theta^*={2\over q}d\zeta^*\biggr\}\;,\end{equation}
with duals
\[\biggr\{\vec e_i={q\over 2}{\partial\over\partial\xi^i}\;;i=1,2\biggr\}
\;\hbox{and}\;
\biggl\{\vec e={q\over 2}{\partial\over\partial\zeta}\;,\;
\vec e^*={q\over 2}{\partial\over\partial\zeta^*}\biggr\}\;.\]
A rotation of the real Zweibein by an angle $\alpha$ :
\[\left(\begin{array}{c}
\theta^1\\ \theta^2\end{array}\right)\Rightarrow
\left(\begin{array}{c}
\tilde\theta^1\\ \tilde\theta^2\end{array}\right)=
\left(\begin{array}{cc}
\cos\alpha & \sin\alpha \\ -\sin\alpha & \cos\alpha\end{array}\right)\;
\left(\begin{array}{c}
\theta^1 \\
\theta^2\end{array}\right)\;,\]
becomes diagonal for the complex Zweibein :
\begin{equation}\label{Hil2}
\left(\begin{array}{c}
\theta\\ \theta^*\end{array}\right)\Rightarrow
\left(\begin{array}{c}
\tilde\theta\\ \tilde\theta^*\end{array}\right)=
\left(\begin{array}{cc}
\exp(-{\bf i}\alpha) & 0 \\ 0 & \exp({\bf i}\alpha)\end{array}\right)\;
\left(\begin{array}{c}
\theta \\ \theta^*\end{array}\right)\;.\end{equation}
This means that the complexified cotangent bundle $(T^*S^2)^{\bf C}$ splits, in an $SO(2)$ invariant way,
into the direct sum of two line bundles $(T^*S^2)^{\prime}$ and $(T^*S^2)^{\prime\prime}$ with
one-dimensional local bases of sections given by $\{\theta\}$ and $\{\theta^*\}$.\\
In the overlap $H_B\cap H_A$, the Zweibein in $H_A$ and in $H_B$ are related by :
\begin{equation}\label{Hil3}
\theta_A=(c_{AB})^{-1}\;\theta_B\;,\;\theta_A^*=c_{AB}\;\theta_B^*\;,\end{equation}
whith the transition function \(c_{AB}=\zeta_B/\zeta_B^*=\zeta_A^*/\zeta_A =\exp(2{\bf i}\varphi)\),
$\varphi$ being the azimuthal angle, well defined (modulo $2\pi$) in $H_B\cap H_A$.\\
Sections of $(T^*S^2)^{\prime}$ and $(T^*S^2)^{\prime\prime}$ are written as
\(\Sigma^{\prime}= \sigma^{(+1)}\;\theta\) and
\(\Sigma^{\prime\prime}=\sigma^{(-1)}\;\theta^*\) such that, in $H_B\cap H_A$,
\[{\sigma^{(\pm 1)}}_{\mid A}=(c_{AB})^{\pm 1}{\sigma^{(\pm 1)}}_{\mid B}\;.\]
Following Staruszkiewicz \cite{Star}, who refers to Pensov, we call such a field a Pensov
scalar of weight $(\pm 1)$. The question is now adressed to define Pensov scalars of
weight $s$ on $S^2$. In general this would require a cocycle condition on transition functions in
triple overlaps. However, since the sphere is covered by only two charts, it is enough that the
overlap equation
\[{\sigma^{(s)}}_{\mid A}=(c_{AB})^s {\sigma^{(s)}}_{\mid B}\]
be well defined. Now, $(c_{AB})^s=\exp(2{\bf i} s\varphi)$ is well defined when $2s$ takes integer
values\footnote{The integer $2s$ can be identified with the integer representing an element of the
second \v{C}ech cohomology group of $S^2$ with integer values, $\check{H}^2(S^2,{\bf Z})={\bf Z}$,
classifying the line bundles over the sphere $S^2$.}.\\ The corresponding line bundle\footnote{In
the sequel, abusing the notation, we shall denote bundles and their spaces of sections by the same
symbol.} will be denoted by ${\cal P}^{(s)}$.\\ A local basis of its sections in $H_B$ is denoted by
${\Theta^{(s)}}_{\mid B}$, and in $H_A$ by ${\Theta^{(s)}}_{\mid A}$. They are related in $H_B\cap
H_A$ by the generalisation of (\ref{Hil3}):
\[{\Theta^{(s)}}_{\mid A}=(c_{AB})^{-s}{\Theta^{(s)}}_{\mid B}\;\]
and a local section is given by :
\(\Sigma^{(s)}={\sigma^{(s)}}\;{\Theta^{(s)}}\;\). \\
On $S^2$ with metric ${\bf g}=\delta_{ij}\theta^i\otimes\theta^j$, the Levi-Civita connection reads
\[\nabla^{LC} \theta^i=-{\bf \Gamma}^i_{\;j}\otimes\theta^j
=-{\Gamma_k^{\;\;i}}_j\;\theta^k\otimes\theta^j\;,\]
where
\[{\Gamma_k^{\;\;i}}_j =
-{1\over 2}\{{\delta^i}_k {\partial q\over\partial\xi^j}-
\delta^{i\ell}{\partial q\over\partial\xi^\ell}\delta_{kj}\}\;.\]
In terms of the complexified Zweibein (\ref{Hil1}), we write :
\begin{equation}\label{Hil4}
\nabla^{LC} \theta=-{\bf \Gamma}\otimes\theta\;,\;
\nabla^{LC} \theta^*=-{\bf \Gamma}^*\otimes\theta^*\;,\end{equation}
where
\[{\bf\Gamma}=-{\bf\Gamma}^*=-{1\over 2}
\{{\partial q\over\partial\zeta}\theta\;-{\partial q\over\partial\zeta^*}\;\theta^*\}=
-{1\over 2}\{\zeta^*\theta-\zeta\theta^*\}\;.\]
It is easy to see that \(\nabla^{LC}\Theta^{(s)}=-s{\bf \Gamma}\otimes\Theta^{(s)}\)
defines a connection in the module of Pensov $s$-scalars generalising (\ref{Hil4}) above.
This connection maps ${\cal P}^{(s)}$ in $(T^*(S^2))^{\bf C}\otimes{\cal P}^{(s)}$. Now the
space of complex-valued one-forms $(T^*(S^2))^{\bf C}$ is isomorphic to ${\cal P}^{(+1)}\oplus{\cal P}^{(-1)}$,
so that $\nabla^{LC}$ is actually a mapping :
\[\nabla^{LC}:{\cal P}^{(s)}\mapsto{\cal P}^{(s+1)}\oplus{\cal P}^{(s-1)}:
\Sigma^{(s)}\mapsto
\nabla^{LC} \Sigma^{(s)}=
\biggl({\bf d}\sigma^{(s)}-s{\bf \Gamma}\sigma^{(s)}\biggr)\otimes\Theta^{(s)}\;.\]
Projecting $\nabla^{LC}\Psi^{(s)}$ on each term in the sum ${\cal P}^{(s+1)}\oplus{\cal P}^{(s-1)}$ we obtain
:
\[\nabla^{LC}\Psi^{(s)} ={1\over 2}\biggl(\check\delta\!\!\!/_s\sigma^{(s)}\Theta^{(s+1)}+
\check\delta\!\!\!/_s^\dagger\sigma^{(s)}\Theta^{(s-1)}\biggr)\;,\]
where we have introduced the "edth" operators of Newman and Penrose \cite{NewPen} :
\begin{eqnarray}
\check\delta\!\!\!/_s\sigma^{(s)}&=&
q^{-s+1}{\partial\over\partial\zeta}\bigl(q^s\sigma^{(s)}\bigr)=
q{\partial\sigma^{(s)}\over\partial\zeta}+
s{\partial q\over\partial\zeta}\sigma^{(s)}\nonumber\;,\\
\check\delta\!\!\!/_s^\dagger\sigma^{(s)}&=&
q^{s+1}{\partial\over\partial\zeta^*}\bigl(q^{-s}\sigma^{(s)}\bigr)=
q{\partial\sigma^{(s)}\over\partial\zeta^*}-
s{\partial q\over\partial\zeta^*}\sigma^{(s)}\;.\label{Hil5}\end{eqnarray}
With respect to the scalar product of Pensov scalars :
\begin{equation}\label{Hil6}
\biggl(\Sigma^{(s)},T^{(s)}\biggr)_s=\int_{S^2}\,\sigma^{(s)*}\,\tau^{(s)}\;\omega\;,\;\end{equation}
where $\omega=\theta^1\wedge\theta^2={{\bf i}\over 2}\theta\wedge\theta^*$ is the invariant volume
element on $S^2$,
the operators $\check\delta\!\!\!/_s$ and $\check\delta\!\!\!/_s^\dagger$ are formally anti-adjoint :
\begin{equation}\label{Hil7}\biggl(\sigma^{(s+1)},\check\delta\!\!\!/_s\tau^{(s)}\biggr)_{s+1}=
\biggl(-\check\delta\!\!\!/_{s+1}^\dagger\sigma^{(s+1)},\tau^{(s)}\biggr)_s\;.\end{equation}
In a previous paper \cite{Migetal1}, the Dirac operator on K\"{a}hler spinors was defined as the
restriction of $-{\bf i}({\bf d}-\delta)$ to the left ideals of the Clifford algebra bundle.
Now, these ideals are identified with ${\cal I}^E_+={\cal P}^{(0)}\oplus{\cal P}^{(+1)}$, with basis
$\{{\bf 1}+{\bf i}\omega,\theta\}$, and ${\cal I}^E_-={\cal P}^{(-1)}\oplus{\cal P}^{(\bar 0)}$, with basis
$\{\theta^*,{\bf 1}-{\bf i}\omega\}$.\\ In these bases, the local expressions of the Dirac operators were
given as :
\begin{eqnarray*}
{\cal D}^E_+\left(\begin{array}{l}\sigma^{(0)}\\ \sigma^{(+1)}\end{array}\right)&=&
-{\bf i}\left(\begin{array}{cc} 0 & \check\delta\!\!\!/^\dagger_{+1}\\ \check\delta\!\!\!/_0 & 0\end{array}\right)
\left(\begin{array}{l}\sigma^{(0)}\\ \sigma^{(+1)} \end{array}\right)\;,\\
{\cal D}^E_-\left(\begin{array}{l}\sigma^{(-1)}\\\sigma^{(\bar 0)}\end{array}\right)&=&
-{\bf i}\left(\begin{array}{cc} 0 & \check\delta\!\!\!/^\dagger_0\\\check\delta\!\!\!/_{-1} & 0\end{array}\right)
\left(\begin{array}{l}\sigma^{(-1)}\\ \sigma^{(\bar 0)}\end{array}\right)\;.\end{eqnarray*}
This suggests to define a Pensov spinor field of weight $s$ as a section
\[\Psi_{(s)}=\Sigma^{(s-1/2)}\oplus\Sigma^{(s+1/2)}\]
of the Whitney sum \({\cal P}^{(s-1/2)}\oplus{\cal P}^{(s+1/2)}\), with a Dirac operator
locally expressed as :
\begin{equation}\label{Hil8}
{\cal D}_{(s)}
\left(\begin{array}{l}
\sigma^{(s-1/2)}\\
\sigma^{(s+1/2)}
\end{array}\right)=
-{\bf i}\left(\begin{array}{cc} 0 & \check\delta\!\!\!/_{s+1/2}^\dagger\\
\check\delta\!\!\!/_{s-1/2} & 0
\end{array}\right)
\left(\begin{array}{l}
\sigma^{(s-1/2)}\\
\sigma^{(s+1/2)}
\end{array}\right)\;.\end{equation}
The usual Dirac spinors on $S^2$ are recovered when $s=0$.\\
With the complex representation of the real Clifford algebra\footnote{
The real Clifford algebra ${\cal C}\ell(p,q)$ is defined by
\(\gamma^k\gamma^\ell+\gamma^\ell\gamma^k=2 \eta^{k\ell}\), where the flat metric tensor
$\eta^{k\ell}$ is diagonal with $p$ times $+1$ and
$q$ times $-1$.
This entails some differences with other work using the Clifford algebra ${\cal C}\ell(0,n)$ for
Riemannian manifolds instead of ${\cal C}\ell(n,0)$ used here.} \({\cal C}\ell(2,0)\)
\begin{equation}\label{gammas}
\gamma^1\Rightarrow\left(\begin{array}{cc} 0 & 1\\ 1 & 0\end{array}\right)\;,\;
\gamma^2\Rightarrow\left(\begin{array}{cc} 0 & {\bf i}\\ -{\bf i} & 0\end{array}\right)\;,\end{equation}
acting on \(\psi_{(s)}=\left(\begin{array}{l}\sigma^{(s-1/2)}\\ \sigma^{(s+1/2)}\end{array}\right)\),
the Dirac operator can also be written as :
\begin{equation}\label{Hil9}
{\cal D}_{(s)}\psi_{(s)} =-{\bf i}\gamma^k\;\nabla_{k,(s)}^{LC}\psi_{(s)}\;,\end{equation}
where the covariant derivative of the spinor \(\psi_{(s)}\) is given by :
\[\nabla_{k,(s)}^{LC}\psi_{(s)}=
{q\over 2}{\partial\psi_{(s)}\over\partial \xi_k}+{1\over 2}
\Sigma_{ij}^{(s)}\Gamma_k^{\;\;ij}\,\psi_{(s)}\;.\] Here, \(\Sigma_{12}^{(s)}=
{\bf i} s\left(\begin{array}{cc}1 & 0\\ 0 & 1\end{array}\right)+
\left(\begin{array}{cc}-{\bf i}/2 & 0\\ 0 & +{\bf i}/2\end{array}\right)\)
reduces to ${1\over 4}[\gamma_1,\gamma_2]$ for $s=0$.\\
In terms of the "edth" operators, we may write
\begin{eqnarray*}
\nabla_{(s),+}^{LC}\psi_{(s)}&=&
{1\over 2}
\left(\begin{array}{cc}\check\delta\!\!\!/_{s-1/2}&0\\ 0&\check\delta\!\!\!/_{s+1/2} \end{array}\right)
\psi_{(s)}\;,\nonumber\\
\nabla_{(s),-}^{LC}\psi_{(s)}&=&
{1\over 2}
\left(\begin{array}{cc}\check\delta\!\!\!/^\dagger_{s-1/2}&0\\
0&\check\delta\!\!\!/^\dagger_{s+1/2}\end{array}\right)\psi_{(s)}\;.\end{eqnarray*}
With \(\gamma^{(+)}=\gamma^1+{\bf i}\gamma^2=
\left(\begin{array}{cc} 0 & 0\\ 2 & 0\end{array}\right)\) and
\(\gamma^{(-)}=\gamma^1-{\bf i}\gamma^2=
\left(\begin{array}{cc} 0 & 2\\ 0 & 0\end{array}\right)\), the Dirac operator of (\ref{Hil8}) is
now written as
\[{\cal D}_{(s)}\psi_{(s)}=-{\bf i}\left(\gamma^{(+)}\nabla_{(s),+}^{LC}
+\gamma^{(-)}\nabla_{(s),-}^{LC}\right)\psi_{(s)}\;.\]
The transformation law for $s$-Pensov fields\footnote{A Pensov spinor of weight $s$ can be
interpreted as a usual Dirac spinor on $S^2$, interacting with a Dirac monopole of strength $s$.
Indeed, in the expression of the covariant derivative, the term ${\bf i} s\gamma^k\Gamma_k^{\;\;12}$ is
the Clifford representative of the one-form (potential)
$\mu_s\doteq{\bf i} s\theta^k\Gamma_k^{\;\;12}=
{s\over 1+\vert\zeta\vert^2}\left(\zeta^*d\zeta-\zeta d\zeta^*\right)$, which, in
$\left\{H_B\;;\;\cos\theta\neq +1\right\}$, takes the usual form ${\mu_s}_{\vert B}={\bf i}
s(1-cos\theta)d\phi$.} under a local Zweibein rotation (\ref{Hil2}) is related to the ${\rm
Spin}^c$ structure of the Pensov spinors :
\[\left(\begin{array}{l}\sigma^{(s-1/2)}\\ \sigma^{(s+1/2)}\end{array}\right)\mapsto
\left(\begin{array}{l}\sigma^{\prime(s-1/2)}\\ \sigma^{\prime(s+1/2)}\end{array}\right)=
\exp\{\alpha\Sigma_{12}^{(s)}\}\left(\begin{array}{l}\sigma^{(s-1/2)}\\
\sigma^{(s+1/2)}\end{array}\right)\;,\]
where
\[\exp\{\alpha\Sigma_{12}^{(s)}\}=
\exp\{{\bf i}\;s\alpha\}
\left(\begin{array}{cc}\exp(-{\bf i}\alpha/2) & 0\\ 0 & \exp(+{\bf i}\alpha/2)\end{array}\right)\;.\]
The Clifford action of $\gamma_3={\bf i}\omega$ yields a grading on the Pensov spinors
\[\gamma_3\left(\begin{array}{l}
\psi^{(s)}\\ \psi^{(s+1)} \end{array}\right)
=\left(\begin{array}{cc} 1 & 0\\ 0 & -1 \end{array}\right)
\left(\begin{array}{l}\psi^{(s-1/2)}\\ \psi^{(s+1/2)}
\end{array}\right)\;,\;(\gamma_3)^2={\bf 1}\;,\]
such that the Dirac operator (\ref{Hil8}) is odd \[{\cal D}_{(s)}\gamma_3\,+\,\gamma_3{\cal D}_{(s)}\;=\;0\;.\]
According to (\ref{Hil6}), the scalar product of two Pensov spinors is defined as :
\begin{equation}\label{Hil10}
\langle\Phi_{(s)}\mid\Psi_{(s)}\rangle=
\biggl(\Sigma^{(s-1/2)},T^{(s-1/2)}\biggr)_{s-1/2}\;+\;
\biggl(\Sigma^{(s+1/2)},T^{(s+1/2)}\biggr)_{s+1/2}\;.\end{equation}
The adjointness (\ref{Hil7}) of $-{\bf i}\check\delta\!\!\!/_{s-1/2}$ and $-{\bf i}\check\delta\!\!\!/^\dagger_{s+1/2}$ implies that
the Dirac operator is formally self-adjoint with respect to this scalar product. After completion,
\({\cal P}^{(s-1/2)}\oplus{\cal P}^{(s+1/2)}\) becomes a bona fide Hilbert space ${\cal H}_{(s)}$ on which
${\cal D}_{(s)}$ acts as a self-adjoint (unbounded) operator. Its spectral resolution is completely
solvable. Indeed, let ${\bf X}=\left(\vec\partial \quad \vec\partial^{\;*}\right)\;\left(\begin{array}{c}X^+\\
X^-\end{array}\right)$ be a vector field on
$S^2$, then the Lie derivatives of the Zweibein along ${\bf X}$ are:
\begin{eqnarray}\label{Hil11}{\cal L}_{{\bf X}}\theta&=&
\biggl({\partial X^+\over\partial\zeta} -{1\over q}({\partial q\over\partial\zeta^*}X^-
+{\partial q\over\partial\zeta}X^+)\biggr)\theta
+{\partial X^+\over\partial\zeta^*}\theta^*\;,\nonumber\\
{\cal L}_{{\bf X}}\theta^*&=&
\biggl({\partial X^-\over\partial\zeta^*} -{1\over q}({\partial q\over\partial\zeta^*}X^-
+{\partial q\over\partial\zeta}X^+)\biggr)\theta^*
+{\partial X^-\over\partial\zeta}\theta
\;.\end{eqnarray}
A vector field ${\bf X}$ is said to be a conformal Killing vector field if
${\cal L}_{{\bf X}}{\bf g}=\mu\,{\bf g}$, where $\mu$ is a scalar function on $S^2$.
The expression of the Lie derivative (\ref{Hil11}) yields then the (anti-)holomorphic constraints :
\[{\partial X^+\over\partial\zeta^*}=0\;,\;{\partial X^-\over\partial\zeta}=0\;,\]
and $\mu$ is given by :
\[\mu=q^2\biggl({\partial (X^+/q^2)\over\partial\zeta}
+{\partial (X^-/q^2)\over\partial\zeta^*}\biggr)\;.\]
If ${\bf X}$ has to be globally defined, its Zweibein components $(2/q)X^+$ and
$(2/q)X^-$ must be finite when $\mid\zeta\mid\rightarrow\infty$.
For the standard metric $q=1+\mid \zeta\mid^2$ and this implies that $X^+$, respectively $X^-$, is
a quadratic polynomial in
$\zeta$, respectively $\zeta^*$.\\
There are thus six linearly independent conformal Killing vector fields, three of which are
genuinely Killing,i.e. with $\mu=0$. They are chosen as\footnote{Here we use the $\zeta_B$
coordinates to conform to standard conventions.}
\begin{eqnarray*}
{\bf i} {\bf L}_x&\;=\;&{{\bf i}\over 2}\biggl((\zeta^2-1)\vec\partial -(\zeta^{*2}-1)\vec\partial^{\;*}\biggr)\;\\
{\bf i} {\bf L}_y&\;=\;&{1\over 2}\biggl((\zeta^2+1)\vec\partial+(\zeta^{*2}+1)\vec\partial^{\;*}\biggr)\;,\\
{\bf i} {\bf L}_z&\;=\;&{\bf i}\biggl(\zeta\vec\partial-\zeta^*\vec\partial^{\;*}\biggr)\;.\end{eqnarray*}
The other three conformal Killing vector fields (with $\mu\neq 0$) are :
\begin{eqnarray*}
{\bf i} {\bf K}_x&\;=\;&{1\over 2}\biggl((\zeta^2-1)\vec\partial+(\zeta^{*2}-1)\vec\partial^{\;*}\biggr)\;,\\
{\bf i} {\bf K}_y&\;=\;&{{\bf i}\over 2}\biggl((-\zeta^2-1)\vec\partial+(\zeta^{*2}+1)\vec\partial^{\;*}\biggr)\;,\\
{\bf i} {\bf K}_z&\;=\;&\biggl(\zeta\vec\partial+\zeta^*\vec\partial^{\;*}\biggr)\;.\end{eqnarray*}
As is well known, they form the Lie algebra $sl(2,{\bf C})$ with its Lie subalgebra
$su(2)$ generated by $\{{\bf i} {\bf L}_x\;,\;{\bf i} {\bf L}_y\;,\;{\bf i} {\bf L}_z\;\}$. Standard angular momentum
technique tells us that is easier to deal with the complex Killing vectors :
\begin{eqnarray}\label{Hil12}
{\bf L}_+&=&{\bf L}_x+{\bf i}{\bf L}_y\;=\;\zeta^2\vec\partial+\vec\partial^{\;*}\;,\nonumber\\
{\bf L}_-&=&{\bf L}_x-{\bf i}{\bf L}_y\;=\;-\vec\partial-\,\zeta^{*2}\vec\partial^{\;*}\;,\nonumber\\
{\bf L}_0&=&{\bf L}_z\;=\;\zeta\vec\partial-\zeta^*\vec\partial^{\;*}\;,\end{eqnarray}
with commutation relations :
\[[{\bf L}_0,{\bf L}_\pm]=\pm\,{\bf L}_\pm\;,\;[{\bf L}_+,{\bf L}_-]=2\,{\bf L}_0\;.\]
The Lie derivatives of $\theta$ with respect to these vector fields are :
\[{\cal L}_+\theta=\zeta\theta\;,\;{\cal L}_-\theta=\zeta^*\theta\;,\;{\cal L}_0\theta=\theta\;.\]
An infinitesimal transformation of the Zweibein by Killing vectors is a rotation of the
form (\ref{Hil2}) :
\(\theta\mapsto\tilde\theta=\theta-{\bf i}\,\delta\alpha\,\theta=\theta+\delta t{\cal L}_{{\bf X}}\theta\;,\)
and, according to (\ref{Hil11}), we identify
\[-{\bf i}\,\delta\alpha=
\biggl({\partial X^+\over\partial\zeta} -{1\over q}(\zeta X^-+\zeta^* X^+)\biggr)\delta t\;.\]
The transformation of the Pensov basis $\Theta^{(s)}$ is then obtained as :
\[\Theta^{(s)}\mapsto\tilde\Theta^{(s)}=\Theta^{(s)}-{\bf i} s\,\delta\alpha\,\Theta^{(s)}=
\Theta^{(s)}+\delta t {\cal L}_{{\bf X}}\Theta^{(s)}\;,\;\hbox{with}\]
\[{\cal L}_{{\bf X}}\Theta^{(s)}=s\biggl({\partial X^+\over\partial\zeta}-{1\over q}
(\zeta X^-+\zeta^* X^+)\biggr)\Theta^{(s)}\;.\]
The Lie derivatives of Pensov fields $\Sigma^{(s)}=\sigma^{(s)}\Theta^{(s)}$ along the the
Killing vectors of (\ref{Hil12}) read :
\begin{eqnarray*}
{\cal L}_+^{(s)}\sigma^{(s)}&\;=\;&
(\zeta^2{\partial\over\partial\zeta}
+{\partial\over\partial\zeta^*}+s\zeta)\sigma^{(s)}\;,\\
{\cal L}_-^{(s)}\sigma^{(s)}&\;=\;&
(-{\partial\over\partial\zeta}
-\zeta^{*2}{\partial\over\partial\zeta^*}+s\zeta^*)\sigma^{(s)}\;,\\
{\cal L}_0^{(s)}\sigma^{(s)}&\;=\;&
(\zeta{\partial\over\partial\zeta}
-\zeta^*{\partial\over\partial\zeta^*}+s)\sigma^{(s)}\;.\end{eqnarray*}
These Lie derivatives yield a representation of $su(2)$ on $s$-Pensov fields :
\[ [{\cal L}_0^{(s)},{\cal L}_\pm^{(s)}]=\pm{\cal L}_\pm^{(s)}\;,
\;[{\cal L}_+^{(s)},{\cal L}_-^{(s)}]=2{\cal L}_0^{(s)}\;.\]
The Casimir operator is given by :
\[(\vec{\cal L}^{(s)})^2={1\over 2}
\biggl({\cal L}_+^{(s)}{\cal L}_-^{(s)}\,+\,{\cal L}_-^{(s)}{\cal L}_+^{(s)}\biggr)+({\cal L}_0^{(s)})^2=
-q^2{\partial^2\over\partial\zeta\partial\zeta^*}+s\,q\,{\cal L}_0^{(s)}\;.\]
Straightforward angular momentum algebra yields the monopole harmonics of Wu and Yang \cite{W-Y},
solutions of
\begin{equation}\label{Hil13} (\vec{\cal L}^{(s)})^2\;{\bf Y}^s_{j,m}=j(j+1)\;{\bf Y}^s_{j,m}\;,\;
{\cal L}_0^{(s)}\;{\bf Y}^s_{j,m}=m\;{\bf Y}^s_{j,m}\;,\end{equation}
where $j=\mid s\mid,\mid s\mid+1,\cdots$ and $m=-j,-j+1,\cdots,j-1,j$.\\
They can be written in terms of Jacobi functions and, using some appropriate Olinde-Rodrigues formulae, they are obtained
as :
\begin{eqnarray}\label{Har}
{\bf Y}^s_{j,m}&=&{(-1)^{j-s}\over 2^j}\sqrt{2j+1\over 4\pi}\sqrt{(j+m)!\over(j+s)!(j-s)!(j-m)!}
\exp({\bf i}(m-s)\varphi)\nonumber\\ &&(1-z)^{-{m-s\over 2}}(1+z)^{-{m+s\over 2}}
\biggl({d\over dz}\biggr)^{j-m}\biggl((1-z)^{j-s}(1+z)^{j+s}\biggr)\;\nonumber\\
&=&{(-1)^{j+m}\over 2^j}\sqrt{2j+1\over 4\pi}\sqrt{(j-m)!\over(j+s)!(j-s)!(j+m)!}
\exp({\bf i}(m-s)\varphi)\nonumber\\ &&(1-z)^{{m-s\over 2}}(1+z)^{{m+s\over 2}}
\biggl({d\over dz}\biggr)^{j+m}\biggl((1-z)^{j+s}(1+z)^{j-s}\biggr)\;,\end{eqnarray}
where $\varphi$ is the azimuthal angle and $z=\cos(\theta)$, the cosine of the polar angle.\\
The Lie derivative of Pensov fields along one of the vector fields
\({\bf L}_\pm,{\bf L}_0\) "commutes" with the edth operators of (\ref{Hil5}) in the sense that\footnote{This is quite expected since the edth operators are constructed from the Levi-Civita connection which is metric compatible and the Killing vector fields conserve this metric.}:
\[\check\delta\!\!\!/_s\;\vec{\cal L}^{(s)}=\vec{\cal L}^{(s+1)}\;\check\delta\!\!\!/_s\;,\;
\check\delta\!\!\!/_{s+1}^\dagger\;\vec{\cal L}^{(s+1)}=\vec{\cal L}^{(s)}\;\check\delta\!\!\!/_{s+1}^\dagger\;.\]
With the choice of phases in (\ref{Har}), one has
\begin{eqnarray*}
\check\delta\!\!\!/_s{\bf Y}^s_{j,m}&=&-\sqrt{(j-s)(j+s+1)}\;{\bf Y}^{s+1}_{j,m}\;,\\
\check\delta\!\!\!/_{s+1}^\dagger{\bf Y}^{s+1}_{j,m}&=&
+\sqrt{(j-s)(j+s+1)}\;{\bf Y}^s_{j,m}\;.\end{eqnarray*}
On Pensov spinors of ${\cal H}_{(s)}$, one defines the "total angular momentum" as :
\[\vec{\cal L}_{tot}^{(s)}\left(\begin{array}{l} \sigma^{(s-1/2)}\\ \sigma^{(s+1/2)}\end{array}\right)
=\left(\begin{array}{cc}
\vec{\cal L}^{(s-1/2)} & 0\\ 0 & \vec{\cal L}^{(s+1/2)}\end{array}\right)
\left(\begin{array}{l} \sigma^{(s-1/2)}\\ \sigma^{(s+1/2)}\end{array}\right)
\;.\]
It commutes with the Dirac operator : \({\cal D}_{(s)}\vec{\cal L}_{tot}^{(s)}\;=\;\vec{\cal L}_{tot}^{(s)}{\cal D}_{(s)}\). \\
From the product of edth operators
\begin{eqnarray*} -\check\delta\!\!\!/_{s+1/2}^\dagger\;\check\delta\!\!\!/_{s-1/2}&=&(\vec{\cal L}^{(s-1/2)})^2-\,(s^2-1/4)\;,\\
-\check\delta\!\!\!/_{s-1/2}\;\check\delta\!\!\!/_{s+1/2}^\dagger&=&(\vec{\cal L}^{(s+1/2)})^2-\,(s^2-1/4)\;,\end{eqnarray*}
a Lichnerowicz type formula follows immediately :
\[{{\cal D}_{(s)}}^2=(\vec{\cal L}_{tot}^{(s)})^2-\;(s^2-1/4)\,{\bf 1}\;.\]
Using (\ref{Hil13}), the eigenvalues of the Dirac operator are found to be :
\[{\cal D}_{(s)}\psi_{(s),j,m}^{(\pm)}=
\pm\sqrt{(j+1/2)^2-s^2}\;\psi_{(s),j,m}^{(\pm)}\;,\]
with
\[\psi_{(s),j,m}^{(+)}=
\left(\begin{array}{c} {1\over\sqrt{2}}{\bf Y}^{s-1/2}_{j,m}\\ {{\bf i}\over\sqrt{2}}{\bf Y}^{s+1/2}_{j,m}
\end{array}\right)\;,\;\;
\psi_{(s),j,m}^{(-)}=
\gamma_3\,\psi^{(+)}_{(s),j,m}=
\left(\begin{array}{c} {1\over\sqrt{2}}{\bf Y}^{s-1/2}_{j,m}\\ -{{\bf i}\over\sqrt{2}}{\bf Y}^{s+1/2}_{j,m}
\end{array}\right)\;.\]
In particular it follows that, for Dirac spinors i.e. $s=0$, and only in this case, there are no zero eigenvalues. When $s\not=0$, zero is an eigenvalue, $2\vert s\vert$ times degenerate with eigenspinors
\[
\left\{\begin{array}{ll}
\left(\begin{array}{c} {\bf Y}^{s-1/2}_{s-1/2,m}\\ 0 \end{array}\right) &
\hbox{ for positive values of s}\;,\\
&\\
\left(\begin{array}{c} 0 \\ {\bf Y}^{s+1/2}_{-s-1/2,m}\end{array}\right) &
\hbox{ if s is negative.}\end{array}\right.\]
\newpage
\section{The projective modules over ${\cal A}$}\label{Module}
\setcounter{equation}{0}
Through the Gel'fand-Na\u{\i}mark construction, the topology of $M=S^2\times\{a,b\}$ is encoded
in the complex $C^*$-algebra of continuous complex-valued functions on $M=S^2\times\{a,b\}$.
However, in order to get a fruitful use of a differential structure, we have to restrict this
$C^*$-algebra to its dense subalgebra of smooth functions. This proviso made, let
$\{f,g,\cdots\}$ denote elements of ${\cal A}={\cal C}(M)$ and let the value of $f$ at a point\footnote{In
this section points of $S^2$ are denoted by $x,y,\cdots$, while $\alpha,\beta,\cdots$ will assume
values in the two-point space $\{a,b\}$. Points of $M=S^2\times\{a,b\}$ are thus written as
$p=\{x,\alpha\},q=\{y,\beta\}$, etc. and the value of a function $F$ at $(p,q,\cdots)$ will also be expressed as
$F_{\alpha,\beta,\cdots}(x,y,\cdots)$.} $p=\{x,\alpha\}\in M$, be written as $f(p)=f_\alpha(x)$.
The vectors of the free right ${\cal A}$-module of rank two, identified with ${\cal A}^2$, are of the form
$X=\sum_{\;i=1,2}E_i\;f^i$, where $f^i\in{\cal A}$ and $\{E_i\,;i=1,2\}$ is a basis of ${\cal A}^2$.
Let $\Omega^\bullet({\cal A})=\sum_{\;k=0}^{\infty}\Omega^{(k)}({\cal A})$ denote the universal differential envelope of ${\cal A}$. Elements of $\Omega^{(k)}({\cal A})$ can be realised, see e.g. \cite{Coq}, as functions on the Cartesian product of $(k+1)$ copies of $M$, vanishing on neighbouring diagonals, i.e.
\(F(p_0,p_1,\cdots,p_k)=0\) if, for some $i$, $p_i=p_{i+1}$.\\
The product in $\Omega^\bullet({\cal A})$ is obtained by concatenation, e.g. if $F\in\Omega^{(1)}({\cal A})$ and
$G\in\Omega^{(2)}({\cal A})$ then their product $F\cdot G\in \Omega^{(3)}({\cal A})$ is represented by
\[(F\cdot G)(p_0,p_1,p_2,p_3)=F(p_0,p_1)\;G(p_1,p_2,p_3)\;.\]
The differential ${\bf d}$ acts on $f\in\Omega^{(0)}({\cal A})$ and on $F\in\Omega^{(1)}({\cal A})$ as follows :
\[({\bf d} f)(p_0,p_1)=f(p_1)-f(p_0)\;,\]
\[({\bf d} F)(p_0,p_1,p_2)=F(p_1,p_2)-F(p_0,p_2)+F(p_0,p_1)\;.\]
The involution\footnote{Note that ${\bf d}(f^\dagger)=-({\bf d} f)^\dagger\;,f\in{\cal A}$ and, more generally, if $F\in\Omega^{(k)}({\cal A})$, then ${\bf d}(F^\dagger)= (-1)^{k+1}({\bf d} F)^\dagger$.}, defined in ${\cal A}$ by $(f^\dagger)(p)=\Bigl(f(p)\Bigr)^*$, extends to $\Omega^{(k)}({\cal A})$ as
\[(F^\dagger)(p_1,p_2,\cdots)=\Bigl(F(\cdots,p_2,p_1)\Bigr)^*\;.\]
A (universal) connection on ${\cal A}^2$ is given, in the basis $\{E_i\;;i=1,2\}$, by an
$\Omega^{(1)}({\cal A})$-valued $2\times 2$ matrix $(\!(\omega)\!)^i_{\;k}$.
It acts on $X=E_i\;f^i$ as :
\begin{equation}\label{PMod1}
\nabla_{free}(X)=E_i\otimes_{\cal A}\Bigl({\bf d} f^i+(\!(\omega)\!)^i_{\;k} f^k\Bigr)\;.\end{equation}
Let $X=E_i\;f^i$ and $Y=E_i\;g^i$ be two vectors of ${\cal A}^2$, then the standard hermitian product
with values in ${\cal A}$ is given by :
\[{\bf h}(X,Y)=\sum_{\;i,j}\;(f^i)^\dagger\;\delta_{\overline{\i}j}g^j=
\sum_{\;i}\;(f^i)^\dagger\;g^i\;\]
It extends as $\Omega^{\bullet}({\cal A})$-valued on
$\Bigl({\cal A}^2\otimes_{\cal A}\Omega^{\bullet}({\cal A})\Bigr)\times
\Bigl({\cal A}^2\otimes_{\cal A}\Omega^{\bullet}({\cal A})\Bigr)$ as
\[{\bf h}(X\otimes F,Y\otimes G)= F^\dagger{\bf h}(X,Y) G\;.\]
The connection is hermitian if \({\bf d}\Bigl({\bf h}(X,Y)\Bigr)=
{\bf h}(X,\nabla_{free} Y)-{\bf h}(\nabla_{free} X,Y)\). \\
For the product above, this yields
\((\!(\omega)\!)^i_{\;j}=
\delta^{i\overline{k}}{(\!(\omega)\!)^\ell_{\;k}}^\dagger\delta_{\overline{\ell}j}\).\\
The representation of $\omega$ by functions on $M\times M$ is given by :
\[(\!(\omega)\!)\Rightarrow\left(
\begin{array}{cc}
K_{\alpha\beta}(x,y) & S_{\alpha\beta}(x,y)\\
T_{\alpha\beta}(x,y) & L_{\alpha\beta}(x,y)\end{array}\right)\]
and the hermiticity condition reads :
\begin{eqnarray}\label{PMod2}
K_{\alpha\beta}(x,y)&=&\Bigl(K_{\beta\alpha}(y,x)\Bigr)^*\;,\nonumber\\
L_{\alpha\beta}(x,y)&=&\Bigl(L_{\beta\alpha}(y,x)\Bigr)^*\:,\nonumber\\
T_{\alpha\beta}(x,y)&=&\Bigl(S_{\beta\alpha}(y,x)\Bigr)^*\;.\end{eqnarray}
The action of the connection (\ref{PMod1}) is represented by :
\[(\nabla_{free} X)^i_{\alpha\beta}(x,y)=
f^i_\beta(y)-f^i_\alpha(x)+{(\!(\omega)\!)^i}_{k,\alpha\beta}(x,y)\;f_\beta^k(y)\;.\]
A projective module is defined by an endomorphism ${\bf P}$ of
${\cal A}^2$ which is idempotent, ${\bf P}^2={\bf P}$, and hermitian, ${\bf P}^\dagger={\bf P}$, where
the adjoint ${\bf A}^\dagger$ of an endomorphism ${\bf A}$ is defined by ${\bf h}(X,{\bf A}^\dagger
Y)={\bf h}({\bf A}X,Y)$. In the basis $\{E_i\}$, the projector is given by a $2\times 2$ matrix
${(P)^i}_j$ with entries in ${\cal A}$ and is represented by ${(P)^i}_{j,\alpha}(x)$. The projective
module ${\cal M}$ is defined as the image of ${\bf P}$ :
\[{\cal M}=\Bigl\{{\bf P} X\;\vert\;X\in {\cal A}^2\Bigr\}=\Bigl\{X\in{\cal A}^2\;\vert\;{\bf P} X=X\Bigr\}\;.\]
The hermiticity of the projector guarantees that ${\bf h}$, restricted to ${\cal M}$, defines a
hermitian product in ${\cal M}$.\\
In the Connes-Lott model \cite{C-L}, the projectors are of the form
\[{(\!( P)\!)^i}_{j,a}(x)={\delta^i}_j\;\hbox{and}\;
{(\!( P)\!)^i}_{j,b}(x)= {1\over 2}\Bigl({\bf 1}+\vec n(x).\vec\sigma\Bigr)^i_{\;j}\;,\]
where $\vec\sigma$ are the Pauli matrices and $\vec n(x)$ is a real unit vector, mapping
$S^2\mapsto S^2$ so that the projectors are classified by $\pi_2(S^2)={\bf Z}$. Furthermore,
since $\pi_2(U(2))=\pi_2(SU(2))=\pi_2(S^3)=\{{\bf 1}\}$, projectors, belonging to
different homoptopy classes, cannot be unitarily equivalent.\\
The target sphere $S^2$ also has two coordinate charts $H_B^{target}$ and $H_A^{target}$.\\
In these charts, the projector $(\!( P_b)\!)$ can be written as
\((\!( P_b^B)\!)=\vert\nu_B\rangle \langle\nu_B\vert\), respectively \((\!(
P_b^A)\!)=\vert\nu_A\rangle \langle\nu_A\vert\), where $\nu_B$, respectively $\nu_A$, is the
complex coordinate of $\vec n$ in $H_B^{target}$, respectively $H_A^{target}$. We have used the
Dirac ket- and bra-notation :
\[\vert\nu_B\rangle={1\over\sqrt{1+\vert\nu_B\vert^2}}
\left(\begin{array}{c} 1\\ \nu_B \end{array}\right)\quad,\quad
\langle \nu_B\vert={1\over\sqrt{1+\vert\nu_B\vert^2}}
\Bigl(1\;\quad\;\nu_B^*\Bigr)\;,\]
\[\vert\nu_A\rangle={1\over\sqrt{1+\vert\nu_A\vert^2}}
\left(\begin{array}{c} -\nu_A \\ 1\end{array}\right)\quad,\quad
\langle \nu_A\vert={1\over\sqrt{1+\vert\nu_A\vert^2}}
\Bigl(-\nu_A^* \;\quad\; 1\Bigr)\;.\]
An element $X$ of ${\cal A}^2$ is represented by the column matrix
\( X\Rightarrow
\left(\begin{array}{c}
\vert f_a(x)\rangle \\ \vert f_b(x)\rangle\end{array}\right)\), where
\(\vert f_a(x)\rangle=\left(\begin{array}{c} f_a^1(x) \\ f_a^2(x)\end{array}\right)\) and
\(\vert f_b(x)\rangle=\left(\begin{array}{c} f_b^1(x) \\ f_b^2(x)\end{array}\right)\).
It belongs to ${\cal M}$ if ${\bf P} X = X$, which yields no restriction on \(\vert f_a(x)\rangle\)
but there is one on \(\vert f_b(x)\rangle\).
In $H_B^{target}$ it is expressed as $\vert f_b\rangle=\vert\nu_B\rangle\;f_b^B$, where
\[f_b^B=\langle\nu_B\vert f_b\rangle
={1\over\sqrt{1+\vert\nu_B\vert^2}}\bigl(f_b^1+\nu_B^*f_b^2\bigr)\;.\]
In the same way, in $H_A^{target}$ one has $\vert f_b\rangle=\vert\nu_A\rangle\;f_b^A\;,$ where
\[f_b^A=\langle\nu_A\vert f_b\rangle
={1\over\sqrt{1+\vert\nu_A\vert^2}}\bigl(-\nu_A^*f_b^1+f_b^2\bigr)\;.\]
As representatives of the homotopy class $[n]\in\pi_2(S^2)\equiv{\bf Z}$, we choose a
mapping $\vec n$ transforming $H_B$, respectively $H_A$ of the range $S^2$, into $H_B^{target}$,
respectively $H_A^{target}$ of the target $S^2$. Such a choice is\footnote{Note that this choice
is different from the one in previous work \cite{Migetal1}.}
\begin{equation}\label{PMod3}
\nu_B(x)=\biggl({\zeta_B\over\zeta_B^*}\biggr)^{n-1\over 2}\zeta_B\;,\;
\nu_A(x)=\biggl({\zeta_A\over\zeta_A^*}\biggr)^{n-1\over 2}\zeta_A\;;\;n\in{\bf Z}\;,\end{equation}
where $\zeta_B\;,\;\zeta_A$ are the complex coordinates of $x\in S^2$.\\
In the overlap $H_A\cap H_B$, with transition function $c_{AB}$ given by (\ref{Hil3}), we have:
\[ f_b^A(x) =\Bigl(c_{AB}(x)\Bigr)^{n/2}\;f_b^B(x)\]
and this tells us that $f_b(x)$ is a Pensov scalar of weight $n/2$.\\
In the rest of this paper we shall omit the $A$ and $B$ labels except when relating quantities
in $H_A$ with those in $H_B$ in the overlap $H_A\cap H_B$.\\
An element of ${\cal M}$ is thus represented by :
\(X\Rightarrow
\left(\begin{array}{c}
\vert f_a\rangle \\\vert \nu\rangle f_b\end{array}\right)\) and its scalar product with
\(Y\Rightarrow
\left(\begin{array}{c}
\vert g_a\rangle \\ \vert \nu\rangle g_b \end{array}\right)\) is given by :
\begin{eqnarray*}
\Bigl({\bf h}_{\bf P}(X,Y)\Bigr)_a(x)&=
&\Bigl(f_a^1(x)\Bigr)^*g_a^1(x)+\Bigl(f_a^2(x)\Bigr)^*g_a^2(x)\;,\\
\Bigl({\bf h}_{\bf P}(X,Y)\Bigr)_b(x)&=
& \Bigl(f_b(x)\Bigr)^*g_b(x)\;.\end{eqnarray*}
An active gauge transformation in the free module ${\cal A}^2$ is given by a unitary $2\times 2$
matrix ${\bf U}$ with values in ${\cal A}$. It retricts to a gauge transformation in ${\cal M}$ when it
commutes with ${\bf P}$ : ${\bf P}{\bf U}={\bf U}{\bf P}$.\\
An element $X\in {\cal M}$ transforms as $X\mapsto {\bf U} X$ given by :
\begin{eqnarray}\label{PMod4}
\vert\Bigl({\bf U} X\Bigr)_a(x)\rangle &=& (\!({\bf U}_a(x))\!)\;\vert f_a(x)\rangle\;,\nonumber\\
\vert\Bigl({\bf U} X\Bigr)_b(x)\rangle &=& \vert\nu(x)\rangle{\bf u}_b(x)\;f_b(x)\;,\end{eqnarray}
where $(\!({\bf U}_a(x))\!)\in U(2)$ and ${\bf u}_b(x)\in U(1)$.\\
The connection in the free module (\ref{PMod1}) induces a connection in the projective
module ${\cal M}$ given by $\nabla X={\bf P}\;\nabla_{free} X$ where $X\in{\cal M}$.\\
It is represented by
\begin{equation}\label{PMod5} \vert(\nabla X)_{\alpha\beta}(x,y)\rangle=
(\!( P_\alpha(x))\!) (\vert f_\beta(y)\rangle
- \vert f_\alpha(x)\rangle)+(\!( A_{\alpha\beta}(x,y))\!)\;\vert f_\beta(y)\rangle\;.\end{equation}
The $2\times 2$ matrices $(\!( A_{\alpha\beta}(x,y))\!)$ are given
by :
\begin{eqnarray}\label{PMod6}
(\!( A_{aa}(x,y))\!)&=&(\!(\omega_{aa}(x,y))\!)\;,\nonumber\\
(\!( A_{ab}(x,y))\!)&=&\vert \Phi_{ab}(x,y)\rangle\langle\nu(y)\vert\;,\nonumber\\
(\!( A_{ba}(x,y))\!)&=&\vert\nu(x)\rangle\,\langle\Phi_{ba}(x,y)\vert\;,\nonumber\\
(\!( A_{bb}(x,y))\!)&=&\vert\nu(x)\rangle\,\omega_b(x,y)\langle\nu(y)\vert\;,\end{eqnarray}
where we have introduced the $\Omega^{(1)}({\cal A})$-valued ket- and bra- vectors :
\begin{eqnarray}\label{PMod7}
\vert\Phi_{ab}(x,y)\rangle&=&(\!(\omega_{ab}(x,y))\!)\vert\nu(y)\rangle\;,\nonumber\\
\langle\Phi_{ba}(x,y)\vert&=&\langle\nu(x)\vert(\!(\omega_{ba}(x,y)\,)\!)\;,\end{eqnarray}
and the universal one-form :
\begin{equation}\label{PMod8}
\omega_b(x,y)=\langle\nu(x)\vert(\!(\omega_{bb}(x,y))\!)\vert\nu(y)\rangle\;.\end{equation}
The hermiticity condition (\ref{PMod2}) yields ;
\begin{eqnarray}\label{PMod9}
(\!(\omega_{aa}(x,y))\!)^+&=&(\!(\omega_{aa}(y,x))\!)\;,\nonumber\\
(\omega_b(x,y))^*&=&\omega_b(y,x)\;,\nonumber\\
\vert\Phi_{ab}(x,y)\rangle^+&=&\langle\Phi_{ba}(y,x)\vert\;.\end{eqnarray}
In $H_B\cap H_A$ :
\begin{eqnarray}\label{PMod10}
\vert\Phi^A_{ab}(x,y)\rangle
&=&\vert\Phi^B_{ab}(x,y)\rangle\,\Bigl(c_{AB}(y)\Bigr)^{-n/2}\;,\nonumber\\
\langle\Phi^A_{ba}(x,y)\vert
&=&\Bigl(c_{AB}(x)\Bigr)^{n/2}\,\langle\Phi^B_{ba}(x,y)\vert\;,\nonumber\\
\omega^A_b(x,y)&=&\Bigl(c_{AB}(x)\Bigr)^{n/2}\,\omega^B_b(x,y)\,\Bigl(c_{AB}(y)\Bigr)^{-n/2}\;.\end{eqnarray}
The action of $\nabla$ on $X\in{\cal M}$ is obtained
using (\ref{PMod6}), (\ref{PMod7}) and (\ref{PMod8}) :
\begin{eqnarray}\label{PMod11}
\vert(\nabla X)_{aa}(x,y)\rangle&=&
\vert f_a(y)\rangle-\vert f_a(x)\rangle
+(\!(\omega_{aa}(x,y))\!)\,\vert f_a(y)\rangle\;,\nonumber\\
\vert(\nabla X)_{ab}(x,y)\rangle&=&
\vert H_{ab}(x,y)\rangle \; f_b(y) - \vert f_a(x)\rangle\;,\nonumber\\
\vert(\nabla X)_{ba}(x,y)\rangle&=&
\vert\nu(x)\rangle\,\biggl[\langle H_{ba}(x,y)\vert f_a(y)\rangle -f_b(x)\biggr]\;,\nonumber\\
\vert(\nabla X)_{bb}(x,y)\rangle&=&
\vert\nu(x)\rangle\,\Bigl[f_b(y)-f_b(x)\nonumber\\
&&\quad\quad\quad\quad +(\omega_b(x,y)+ m_b(x,y))f_b(y)\Bigr]\;,
\end{eqnarray}
where
\begin{eqnarray}\label{PMod12}
\vert H_{ab}(x,y)\rangle&=&\vert\Phi_{ab}(x,y)\rangle+\vert\nu(y)\rangle\;,\nonumber\\
\langle H_{ba}(x,y)\vert&=&\langle\Phi_{ba}(x,y)\vert+\langle\nu(x)\vert\;.\end{eqnarray}
and the "monopole" connection $m_b(x,y)$ appears as :
\begin{equation}\label{PMod13}
m_b(x,y)=\langle\nu(x)\vert\nu(y)\rangle\,-1\;.\end{equation}
As seen from (\ref{PMod10}) and (\ref{PMod12}), the off-diagonal connections
$\vert H_{ab}(x,y)\rangle$, $\langle H_{ba}(x,y)\vert$ and also $\omega_b(x,y)$ transform
homogeneously from $H_B$ to $H_A$ but $m_b(x,y)$ transforms with the expected inhomogeneous term :
\begin{eqnarray}\label{PMod14}
m^A_b(x,y)&=&
(c_{AB}(x))^{n/2}\,m^B_b(x,y)(c_{AB}(y))^{-n/2}\nonumber\\
&& +(c_{AB}(x))^{n/2}[(c_{AB}(y))^{-n/2}-(c_{AB}(x))^{-n/2}]\;.\end{eqnarray}
In terms of abstract universal differential one forms, (\ref{PMod13}) and (\ref{PMod14}) read :
\begin{eqnarray*}
m_b&=&{1\over\sqrt{1+\vert\nu\vert^2}}\Bigl(\nu^*{\bf d}\nu - \sqrt{1+\vert\nu\vert^2}
{\bf d} \sqrt{1+\vert\nu\vert^2}\Bigr){1\over\sqrt{1+\vert\nu\vert^2}}\;,\\
m_b^A&=&(c_{AB})^{n/2}\,m_b^B\,(c_{AB})^{-n/2}+(c_{AB})^{n/2}{\bf d}(c_{AB})^{-n/2}\;.\end{eqnarray*}
The curvature of the connection is defined by :
\(\vert{\nabla}^2X\rangle = (\!( R)\!)\vert X\rangle\). It is a right-module
homomorphism ${\cal M}\rightarrow{\cal M}\otimes_{\cal A}\Omega^{(2)}({\cal A})$ given
in the basis$\{E_i\}$ by the $2\times 2$ matrix with values in $\Omega^{(2)}({\cal A})$ :
\[(\!( R )\!)= (\!( P)\!) \Bigl({\bf d}(\!( A)\!)\Bigr) (\!( P)\!)+ (\!( A)\!)^2 +
(\!( P)\!) \Bigl({\bf d} (\!( P)\!)\Bigr)\Bigl( {\bf d} (\!( P)\!)\Bigr) (\!( P)\!)\;,\]
or, within the used realisation, by
\begin{eqnarray}\label{PMod15}
\lefteqn{(\!( R_{\alpha\beta\gamma}(x,y,z))\!)=}\hspace{5ex}\nonumber \\
&&(\!( P_\alpha(x))\!)
\Bigl((\!( A_{\beta\gamma}(y,z))\!)-(\!( A_{\alpha\gamma}(x,z))\!)+
(\!( A_{\alpha\beta}(x,y))\!)\Bigr)(\!( P_\beta(z))\!)\nonumber\\
&& + (\!( A_{\alpha\beta}(x,y))\!)\,(\!( A_{\beta\gamma}(y,z))\!)\nonumber\\
&& +(\!( P_\alpha(x))\!) \Bigl((\!( P_\beta(y))\!)-(\!( P_\alpha(x))\!)\Bigr)
\Bigl((\!( P_\gamma(z))\!)-(\!( P_\beta(y))\!)\Bigr)(\!( P_\gamma(z))\!)\;.\nonumber\\
&&\end{eqnarray}
A connection $\nabla$ compatible with the hermitian structure in ${\cal M}$ implies in a
self-adjoint curvature :
\begin{equation}\label{PMod16}
{R^i}_j=\delta^{i\bar\ell} ({R^k}_\ell)^+\delta_{\bar k j}\;.\end{equation}
Let the connection $\nabla$ be extended to ${\cal M}\otimes_{\cal A}\Omega^\bullet({\cal A})$ by
\[\nabla\Bigl(X\otimes_{\cal A} F\Bigr)=(\nabla X)F+X\otimes_{\cal A} {\bf d} F\;,\]
then $\nabla^2$ becomes an endomorphism of the right $\Omega^\bullet({\cal A})$-module
${\cal M}\otimes_{\cal A}\Omega^\bullet({\cal A})$.\\
The active gauge transformation (\ref{PMod4}) acts on the right on the space of connections
as $\nabla\mapsto\nabla^{\bf U}={\bf U}^{-1}\circ\nabla\circ{\bf U}$. The action of $\nabla^{\bf U}$ on $X$ is
given by a similar expression as in (\ref{PMod5}) with the matrices $(\!( A)\!)$ replaced
by $(\!( A^{\bf U})\!)=(\!( {\bf U})\!)^{-1}(\!( A)\!) (\!({\bf U})\!)+
(\!( P)\!)(\!({\bf U}^{-1})\!)({\bf d}(\!({\bf U})\!))(\!( P)\!)$ and (\ref{PMod11}) becomes
\begin{eqnarray*}
\vert(\nabla^{\bf U} X)_{aa}(x,y)\rangle&=&
\vert f_a(y)\rangle - \vert f_a(x)\rangle + (\!(\omega_{aa}^{\bf U}(x,y))\!)\,\vert f_a(y)\rangle\;,\\
\vert(\nabla^{\bf U} X)_{ab}(x,y)\rangle&=&
\vert H_{ab}^{{\bf U}}(x,y)\rangle f_b(y) - \vert f_a(x)\rangle\;,\\
\vert(\nabla^{\bf U} X)_{ba}(x,y)\rangle&=&
\vert\nu(x)\rangle\,\biggl[\langle H_{ba}^{{\bf U}}(x,y)\vert f_a(y)\rangle-f_b(x)\biggr]\;,\\
\vert(\nabla^{\bf U} X)_{bb}(x,y)\rangle&=&
\vert\nu(x)\rangle\,\Bigl[f_b(y)-f_b(x)\nonumber\\
&&\quad\quad\quad\quad+(\omega_b^{{\bf U}}(x,y)+m_b^{{\bf U}}(x,y))f_b(y)\Bigr]\;.
\end{eqnarray*}
with
\begin{eqnarray}\label{PMod17}
(\!(\omega_{aa}^{\bf U}(x,y))\!)&=&(\!({\bf U}_a(x))\!)^{-1}
(\!(\omega_{aa}(x,y))\!)(\!({\bf U}_a(y))\!)\nonumber\\
&&+(\!(({\bf U}_a(x))\!)^{-1}\Bigl((\!({\bf U}_a(y))\!)-(\!({\bf U}_a(x))\!)\Bigr)\;,\nonumber\\
\vert H_{ab}^{{\bf U}}(x,y)\rangle&=&
(\!({\bf U}_a(x))\!)^{-1}\;\vert H_{ab}(x,y)\rangle\;{\bf u}_b(y)\;,\nonumber\\
\langle H_{ba}^{{\bf U}}(x,y)\vert&=&
({\bf u}_b(x))^{-1}\;\langle H_{ba}(x,y)\vert\;(\!({\bf U}_a(y))\!)\;,\nonumber\\
m_b^{{\bf U}}(x,y)&=&
({\bf u}_b(x))^{-1}\;m_b(x,y)\;{\bf u}_b(y)\nonumber\\
\omega_b^{{\bf U}}(x,y)&=&
({\bf u}_b(x))^{-1}\;\omega_b(x,y)\;{\bf u}_b(y)\nonumber\\
&&+({\bf u}_b(x))^{-1}\Bigl({\bf u}_b(y)-{\bf u}_b(x)\Bigr)\;.\end{eqnarray}
It is thus seen that $\vert H_{ab}(x,y)\rangle\,,\,\langle H_{ba}(x,y)\vert$ and
the monopole connection (\ref{PMod13}) $m_b(x,y)$ transform homogeneously under an
active gauge transformation, while $(\!(\omega_{aa}(x,y))\!)$ and $\omega_b(x,y)$ have
the expected inhomogeneous terms ${\bf U}^{-1}{\bf d}{\bf U}$, ${\bf u}^{-1}{\bf d}{\bf u}$.
\newpage
\section{The spectral triple $\Bigl\{{\cal A},{\cal H},{\bf D},\Gamma\Bigr\}$}\label{Triple}
\setcounter{equation}{0}
An (even) spectral triple,$\{{\cal A},{\cal H},{\bf D},\chi\}$, as defined by Connes \cite{Con1}, is given by
a $C^\star$ algebra ${\cal A}$ and a Hilbert space ${\cal H}$, graded by $\chi$, with
\(\pi:{\cal A}\mapsto{\cal B}({\cal H}): f\mapsto \pi(f)\), a faithful, ${\rm Ker}(\pi)=0$, and even,
\(\Bigl[\chi,\pi(f)\Bigr]=0\), $\star$-representation, \(\pi(f^*)=\Bigl(\pi(f)\bigr)^+\), of ${\cal A}$
acts as bounded operators.\\ Furthermore, on ${\cal H}$, there is a self-adjoint Dirac operator
${\bf D}$, which is odd, \({\bf D}\chi+\chi{\bf D}=0$, and such that $({\bf D}-\lambda)^{-1}$ is
compact for $\lambda \not\in {\bf R}$. It should be a first order operator in the sense that
\(\Bigl[[{\bf D},\pi(f)],\pi(g)\Bigr]=0\;,\;f,g\in{\cal A}\).\\ Here the algebra is
${\cal A}={\cal C}(S^2\times\{a,b\};{\bf C})$ and as Hilbert space we take
\[{\cal H}={\cal H}_{(s)}\otimes{\cal H}_{dis}\;,\]
the tensor product of the Hilbert space ${\cal H}_{(s)}$ of Pensov spinors with a finite Hilbert space
\({\cal H}_{dis}=\Bigl({\bf C}^{N_a}\oplus{\bf C}^{N_b}\Bigr)\) where
$N_a,N_b$ are natural numbers giving the number of generations in each chirality sector.\\
\({\cal H}_{dis}\) is endowed with a grading operator:
\begin{equation}\label{TRIP0}\chi_{dis}=\left(\begin{array}{cc}
{\bf 1}_a & 0 \\ 0 & -{\bf 1}_b\end{array}\right)\;,\end{equation}
where ${\bf 1}_\alpha\;,\;\alpha=a,b$ is the $N_\alpha\times N_\alpha$ unit matrix.
On ${\cal H}_{dis}$ there a finite Dirac operator ${\bf D}_{dis}$ represented by a hermitian matrix
odd with respect to $\chi_{dis}$ :
\begin{equation}\label{TRIP1}
{\cal D}_{dis}=\left(\begin{array}{cc}0 & M^+\\ M & 0\end{array}\right)\;,\end{equation}
where $M$ is a $N_b\times N_a$ matrix describing the phenomenology of the
masses.
The total grading in \({\cal H}\) is given by
\begin{equation}\label{TRIP01}\chi=\gamma_3\otimes\chi_{dis}=\left(\begin{array}{cc}
\gamma_3\otimes{\bf 1}_a & 0 \\ 0 &
-\gamma_3\otimes{\bf 1}_b\end{array}\right)\;,\end{equation}
The Dirac operator in ${\cal H}$ is obtained from (\ref{Hil9}) and (\ref{TRIP1}) as ;
\begin{equation}\label{TRIP2}
{\bf D}={\bf D}_{(s)}\otimes{\bf 1}_{a+b}+\gamma_3\otimes{\bf D}_{dis}\;.\end{equation}
It is odd with respect to the total grading $\chi$. The grading operator $\gamma_3$ has
been introduced in (\ref{TRIP2}) so that the square of the total Dirac operator reads
\[{\bf D}^2
={{\bf D}_{(s)}}^2\otimes{\bf 1}_{a+b}+{\bf 1}\otimes{{\bf D}_{dis}}^2\;.\]
An element ${\bf \Psi}$ of ${\cal H}$ is represented as
\begin{equation}\label{TRIP3}
\left(\begin{array}{c}\psi_{(s),a}(x)\\ \psi_{(s),b}(x)\end{array}\right)\;,\end{equation}
where each $\psi_{(s),\;\alpha}(x),(\alpha=a,b)$, is a Pensov spinor with
$N_\alpha$ "generation" indices\footnote{These "generation" indices are not written down
explicitely and the $(s)$ subscript, fixed once for all, will also be omitted in this section.}.
The scalar product is the obvious extension of (\ref{Hil10}) :
\begin{equation}\label{TRIP4}
({\bf \Psi};{\bf \Phi})=\int_{S^2}\biggl[(\psi_a)^+\phi_a\,+\,(\psi_b)^+\phi_b\biggr]\omega\;.\end{equation}
An element $f\in{\cal A}$ is represented as:
\begin{equation}\label{TRIP5}
\pi(f)\;\left(\begin{array}{c}\psi_a(x)\\ \psi_b(x)\end{array}\right)=
\left(\begin{array}{cc}
f_a(x){\bf 1}_c\otimes{\bf 1}_a & 0 \\
0 & f_b(x){\bf 1}_c\otimes{\bf 1}_b
\end{array}\right)\;\left(\begin{array}{c}\psi_a(x)\\ \psi_b(x)\end{array}\right)\;,\end{equation}
where ${\bf 1}_c$ is the $2\times 2$ unit matrix in the Clifford algebra.
Acting on vectors of the form (\ref{TRIP3}), (\ref{TRIP2}) becomes
\[{\cal D}=\left(\begin{array}{cc}
{\cal D}_{(s)}\otimes{\bf 1}_a & \gamma_3\otimes M^+\\
\gamma_3\otimes M & {\cal D}_{(s)}\otimes {\bf 1}_b
\end{array}\right)\;.\]
Its commutator of with $\pi(f)$ is :
\begin{equation}\label{TRIP6}
\bigl[{\cal D},\pi(f)\bigr]=
\left(\begin{array}{cc}
-{\bf i}\;{\bf c}({\rm d}f_a)\otimes{\bf 1}_a & (f_b-f_a)\gamma_3\otimes M^+ \\
(f_a-f_b)\gamma_3\otimes M & -{\bf i}\;{\bf c}({\rm d}f_b)\otimes{\bf 1}_b
\end{array}\right)\;,\end{equation}
where the de Rham exterior differential is
\({\rm d}f_a=\Bigl(\vec e_k(f_a)\Bigr)\theta^k\) and where ${\bf c}(\sigma^{(k)})$ denotes
the Clifford representation of the k-form $\sigma^{(k)}$ :
\[{\bf c}(\sigma_{i_1...i_k}\;\theta^{i_1}\wedge...\wedge\theta^{i_k})=
\sigma_{i_1...i_k}\;\gamma^{i_1}...\gamma^{i_k}\;.\]
The representation $\pi$ of (\ref{TRIP5}) extends to a $\star$-representation of
$\Omega^\bullet({\cal A})$ by :
\[\pi(f_0{\bf d} f_1\cdots{\bf d} f_k)=
\pi(f_0)[{\bf D},\pi(f_1)]\cdots[{\bf D},\pi(f_k)]\;.\]
From (\ref{TRIP5}) and (\ref{TRIP6}) it follows that the element
$f{\bf d} g\in \Omega^{(1)}({\cal A})$ is represented by
\begin{equation}\label{TRIP7}
\pi(f{\bf d} g)=
\left(\begin{array}{cc}
f_a-{\bf i}\;{\bf c}({\rm d}g_a)\otimes{\bf 1}_a
& f_a(g_b-g_a)\gamma_3\otimes M^+ \\
f_b(g_a-g_b)\gamma_3\otimes M &
f_b-{\bf i}\;{\bf c}({\rm d}g_b)\otimes{\bf 1}_b
\end{array}\right)\;,\end{equation}
A general element $F\in \Omega^{(1)}({\cal A})$, given by $F_{\alpha\;\beta}(x,y)$, is then
represented as an operator on ${\cal H}$ by :
\begin{equation}\label{TRIP8}
\pi(F)=
\left(\begin{array}{cc}
-{\bf i}\;{\bf c}(\sigma^{(1)}_a)\otimes{\bf 1}_a &
\sigma^{(0)}_{ab}\;\gamma_3\otimes M^+ \\
\sigma^{(0)}_{ba}\;\gamma_3\otimes M &
-{\bf i}\;{\bf c}(\sigma^{(1)}_b)\otimes{\bf 1}_b
\end{array}\right)\;,\end{equation}
where the $\sigma^{(k)}$'s are differential k-forms given by :
\begin{equation}\label{one-form}
\sigma_a^{(1)}(x)=\Bigl(\vec e_{k,y}\,F_{aa}(x,y)\Bigr)_{\mid y=x}\;\theta^k_x\;,\;
\sigma_b^{(1)}(x)=\Bigl(\vec e_{k,y}\,F_{bb}(x,y)\Bigr)_{\mid y=x}\;\theta^k_x\,,\]
\[\sigma_{ab}^{(0)}(x)=F_{ab}(x,y)_{\mid y=x}\;,\;
\sigma_{ba}^{(0)}(x)=F_{ba}(x,y)_{\mid y=x}\;.\end{equation}
The representative of a universal 2-form $f{\bf d} g{\bf d} h$ will be given by the product of
the matrix (\ref{TRIP7}) with
\[\left(\begin{array}{cc} -{\bf i}\;{\bf c}({\rm d}h_a)\otimes{\bf 1}_a
& (h_b-h_a)\gamma_3\otimes M^+ \\ (h_a-h_b)\gamma_3\otimes M &
-{\bf i}\;{\bf c}({\rm d}h_b)\otimes{\bf 1}_b \end{array}\right)\;.\]
The result is :
\[\pi(f{\bf d} g{\bf d} h)=\left(\begin{array}{cc}
\pi(f{\bf d} g{\bf d} h)_{[aa]} & \pi(f{\bf d} g{\bf d} h)_{[ab]}\\
\pi(f{\bf d} g{\bf d} h)_{[ba]} & \pi(f{\bf d} g{\bf d} h)_{[bb]}\end{array}\right)\;,\]
where
\[\pi(f{\bf d} g{\bf d} h)_{[aa]}=
-f_a\;{\bf c}({\rm d}g_a)\;{\bf c}({\rm d}h_a)\otimes{\bf 1}_a
+f_a(g_b-g_a)(h_a-h_b)\otimes M^+M \;,\]
\[\pi(f{\bf d} g{\bf d} h)_{[ab]}=
-{\bf i}\;\Bigl(f_a\;{\bf c}({\rm d}g_a)(h_b-h_a)
- f_a(g_b-g_a)\;{\bf c}({\rm d}h_b)\Bigr)\gamma_3\otimes M^+ \;,\]
\[\pi(f{\bf d} g{\bf d} h)_{[ba]}=
-{\bf i}\;\Bigl(-f_b(g_a-g_b)\;{\bf c}({\rm d}h_a)
+f_b\;{\bf c}({\rm d}g_b)(h_a-h_b)\Bigr)\gamma_3\otimes M\;,\]
\[\pi(f{\bf d} g{\bf d} h)_{[bb]}=
-f_b\;{\bf c}({\rm d}g_b)\;{\bf c}({\rm d}h_b)\otimes{\bf 1}_b
+f_b(g_a-g_b)(h_b-h_a)\otimes MM^+\;.\]
A generic universal two-form $G$ is represented by
\begin{equation}\label{TRIP9}
\pi(G)=\left(\begin{array}{cc}
\pi(G)_{[aa]} & \pi(G)_{[ab]}\\
\pi(G)_{[ba]} & \pi(G)_{[bb]}\end{array}\right)\;,\end{equation}
with\footnote{Here we have used ${\bf c}(\theta^k){\bf c}(\theta^\ell)=
{\bf c}(\theta^k\wedge\theta^\ell) + \delta^{k\ell}$.}
\begin{eqnarray*}
\pi(G)_{[aa]}&=&-{\bf c}(\rho_{aaa}^{(2+0)})\otimes{\bf 1}_a+
\rho_{aba}^{(0)}\otimes M^+M \;,\\
\pi(G)_{[ab]}&=&-{\bf i}\;{\bf c}(\rho_{ab}^{(1)})\gamma_3\otimes M^+\;,\\
\pi(G)_{[ba]}&=&-{\bf i}\;{\bf c}(\rho_{ba}^{(1)})\gamma_3\otimes M\;,\\
\pi(G)_{[bb]}&=&-{\bf c}(\rho_{bbb}^{(2+0)})\otimes{\bf 1}_b+
\rho_{bab}^{(0)}\otimes MM^+\;,
\end{eqnarray*}
where the differential forms $\rho^{(k)}(x)$ are given by :
\begin{eqnarray}
\rho_{aaa}^{(2+0)}(x)&=&\rho_{aaa}^{(2)}+\rho_{aaa}^{(0)}\;,\nonumber\\
\rho_{bbb}^{(2+0)}(x)&=&\rho_{bbb}^{(2)}+\rho_{bbb}^{(0)}\;,\nonumber\\
\rho_{aaa}^{(2)}(x)&=&
\Bigl({1\over 2}[\vec e_{k,y}\vec e_{\ell,z}-\vec e_{\ell,y}\vec e_{k,z}]G_{aaa}(x,y,z)\Bigr)_{\mid y=x,z=x}\,
\theta_x^k\wedge\theta_x^\ell\;,\nonumber\\
\rho_{bbb}^{(2)}(x)&=&
\Bigl({1\over 2}[\vec e_{k,y}\vec e_{\ell,z}-\vec e_{\ell,y}\vec e_{k,z}]G_{bbb}(x,y,z)\Bigr)_{\mid y=x,z=x}\,
\theta_x^k\wedge\theta_x^\ell\;,\nonumber\\
\rho_{aaa}^{(0)}(x)&=&\delta^{kl}\,
\Bigl(\vec e_{k,y}\vec e_{\ell,z}G_{aaa}(x,y,z)\Bigr)_{\mid y=x,z=x}\;,\nonumber\\
\tilde\rho_{bbb}^{(0)}(x)&=&\delta^{kl}\,
\Bigl(\vec e_{k,y}\vec e_{\ell,z}G_{bbb}(x,y,z)\Bigr)_{\mid y=x,z=x}\;,\nonumber\\
\rho_{aba}^{(0)}(x)&=&G_{aba}(x,y,z)_{\mid y=x,z=x}\;,\nonumber\\
\rho_{bab}^{(0)}(x)&=&G_{bab}(x,y,z)_{\mid y=x,z=x}\;,\nonumber\\
\rho_{ab}^{(1)}(x)&=&
\Bigl(\vec e_{k,y}G_{aab}(x,y,z)-
\vec e_{k,z}G_{abb}(x,y,z)\Bigr)_{\mid y=x,z=x}\,\theta_x^k\;,\nonumber\\
\rho_{ba}^{(1)}(x)&=&
\Bigl(\vec e_{k,y}G_{bba}(x,y,z)-
\vec e_{k,z}G_{baa}(x,y,z)\Bigr)_{\mid y=x,z=x}\,\theta_x^k\;.\label{two-form}
\end{eqnarray}
The representation $\pi$ of $\Omega^\bullet({\cal A})$ is a $\star$-representation but it is not a
differential representation of $\Omega^\bullet({\cal A})$ in the sense that
$G\in Ker(\pi)={\cal J}_0$ does not imply $\pi({\bf d} G)=0$.
To obtain a graded differential algebra of operators in ${\cal H}$, it is necessary to take the quotient
of $\Omega^\bullet({\cal A})$ by the graded differential ideal ${\cal J}={\cal J}_0+{\bf d}{\cal J}_0$
with canonical projection
\begin{equation}\label{TRIP10}
\pi_D:\Omega^\bullet({\cal A})\mapsto \Omega^\bullet_D({\cal A})={\Omega^\bullet({\cal A})\over{\cal J}}=
\bigoplus_{k=0}^\infty\;\Omega_D^{(k)}({\cal A})\;.\end{equation}
Using the homomorphism theorem for algebras, it is easily seen that
\[\Omega_D^{(k)}({\cal A})={\pi\Bigl(\Omega^{(k)}({\cal A})\Bigr)\over\pi({\bf d}{\cal J}_0^{(k-1)})}\;.\]
Since $\pi$ is a faithful representation of ${\cal A}$, its kernel ${\cal J}_0^{(0)}$ has to be zero and
\(\Omega^{(1)}_D({\cal A})\cong\pi\Bigl(\Omega^{(1)}({\cal A})\Bigr)\). So $F\in\Omega^{(1)}({\cal A})$
has in $\Omega_D^{(1)}({\cal A})$ the same representative as given by (\ref{TRIP8}).\\
To compute $\Omega^{(2)}_D({\cal A})$, we need ${\bf d}{\cal J}_0^{(1)}$. According to (\ref{TRIP8}),
${\cal J}^{(1)}_0$ is given by :
\[\Bigl\{F\in \Omega^{(1)}({\cal A})\Vert\;F_{ab}(x,x)=0=F_{ba}(x,x)\,;\,
{\vec e_{k,y}\,F_{\alpha\alpha}(x,y)}_{\mid y=x}=0,\alpha=a,b\Bigr\}\]
so that (\ref{TRIP9}) with
$G={\bf d} F$ and $F\in {\cal J}_0^{(1)}$, yield a \(\pi({\bf d} F)\) of the form
\[\left(\begin{array}{cc}j_a^{(0)}\otimes{\bf 1}_a & 0 \\
0 & j_b^{(0)}\otimes{\bf 1}_b\end{array}\right)\;,\]
where \(j_\alpha^{(0)}(x)=
-\delta^{k\ell}\Bigl(\vec e_{k,y}\vec e_{\ell,z}F_{\alpha\alpha}(y,z)\Bigr)_{\mid y=x,z=x}\),
with \(\alpha=a,b\). \\
In the space \(\pi\Bigl(\Omega^{(k)}({\cal A})\Bigr)\), whose elements are bounded operators in
${\cal H}$, the scalar product of $\pi(G_1)$ and $\pi(G_2)$ is defined by
\[\langle \pi(G_1);\pi(G_2)\rangle_k =
{\bf Tr}_{Dix}\Bigl\{{\pi(G_1)}^+\pi(G_2)\mid{\bf D}\mid^{-d}\Bigr\}\;,\]
where ${\bf Tr}_{Dix}$ is the Dixmier trace and $d$ (here $d=2$) is the dimension
of the spectral triple as defined in Connes' book \cite{Con1} .\\
With respect to this scalar product, \(\pi\Bigl(\Omega^{(k)}({\cal A})\Bigr)\) can be completed to a
Hilbert space ${\cal H}^{(k)}$ and its quotient by \(\pi({\bf d}{\cal J}_0^{(k-1)})\), i.e.
\(\Omega^{(k)}_D({\cal A})\), will be a dense subspace of \(\pi({\bf d}{\cal J}_0^{(k-1)})^\perp\),
the orthogonal complement of $\pi\Bigl({\bf d}{\cal J}_0^{(k-1)}\Bigr)$.\\
Let ${\cal P}^{(k)}$ be the projector on \({\cal H}^{(k)}_D=\pi({\bf d}{\cal J}_0^{(k-1)})^\perp\),
then a scalar product in \(\Omega^{(k)}_D({\cal A})\) is defined by :
\begin{equation}\label{TRIP11}
\langle \pi_D(G_1);\pi_D(G_2)\rangle_{k,D}=
\langle {\cal P}^{(k)}(\pi(G_1));{\cal P}^{(k)}(\pi(G_2))\rangle_k\;.\end{equation}
It can also be shown that \({\cal H}^{(k)}_D\) is a Hilbert ${\cal A}$-bimodule with a two-sided
representation of the unitaries \(\;{\cal U}({\cal A})=\{u\in {\cal A}\vert uu^+=u^+u=1\}\).
Indeed, if $G\in\Omega^{(k)}({\cal A})$ then $uG$ and $Gu$ also belong to $\Omega^{(k)}({\cal A})$ so that
${\cal H}^{(k)}$ is a Hilbert ${\cal A}$-bimodule. Furthermore $\pi(u)\pi_D(G)=\pi_D(uG)$
and $\pi_D(G)\pi(u)=\pi_D(G u)$ show that ${\cal P}^{(k)}:{\cal H}^{(k)}\rightarrow{\cal H}^{(k)}_D$
is a bimodule homomorphism and the Dixmier trace properties guarantee that
\begin{eqnarray}\label{TRIP12}
\langle \pi_D(G_1);\pi_D(G_2)\rangle_{k,D}&=&
\langle\pi(u)\pi_D(G_1);\pi(u)\pi_D(G_2)\rangle_{k,D}\nonumber\\
&=&\langle \pi_D(G_1)\pi(u);\pi_D(G_2)\pi(u)\rangle_{k,D}\;.\end{eqnarray}
The following trace theorems of Connes \cite{Con1} will be needed in the sequel of the calculations.
{\bf 1)} If $A_s\otimes B$ is a bounded operator in ${\cal H}={\cal H}_{(s)}\otimes{\bf C}^N$, then
\begin{equation}\label{TRIP13}
{\bf Tr}_{Dix}\biggl\{(A_s\otimes B)\mid{\bf D}\mid^{-2}\biggr\}=
{\bf Tr}_{Dix}\Bigl\{A_s\mid{\cal D}_{(s)}\mid^{-2}\Bigr\}\;{\bf tr}\{B\}\;,\end{equation}
where {\bf tr} is the ordinary trace on $N\times N$ matrices.
{\bf 2)} Let $A_s$ be a section of the Clifford bundle over a compact d-dimensional
manifold $M\;(=S^2)$ with its action on the (Pensov) spinors, then
\begin{equation}\label{TRIP14}
{\bf Tr}_{Dix}\Bigl\{A_s\mid{\cal D}_{(s)}\mid^{-d}\Bigr\}=
({1\over 4\pi})^{d/2}{1\over \Gamma(d/2+1)}\int_M\,{\bf tr}_{{\bf c}}\{A_s\}\omega\;,\end{equation}
where ${\bf tr}_{{\bf c}}$ is the trace on the representation of the Clifford algebra.\\
\vspace{0.3cm}
\noindent
An element $\pi(G)$ of $\Omega_D^{(2)}({\cal A})\cong \Bigl(\pi({\bf d}{\cal J}_0^{(1)})\Bigr)^\perp$
is of the form given in (\ref{TRIP9}) and has to obey :
\[\forall\;j_\alpha^{(0)}\,:{\bf Tr}_{Dix}\biggl\{\left(\begin{array}{cc}
j_a^{(0)}(x)^*\otimes{\bf 1}_a & 0 \\0 & j_b^{(0)}(x)^*\otimes{\bf 1}_b\end{array}\right)
\pi(G)\mid{\bf D}\mid^{-2}\biggl\}=0\;.\]
Using the trace properties (\ref{TRIP13}) and (\ref{TRIP14}), we obtain the orthogonality condition :
\begin{equation}\label{Nojunk}
\rho_{aaa}^{(0)}-{1\over N_a}\rho_{aba}^{(0)}\;{\bf tr}\{M^+M\}=0\quad;\quad
\rho_{bbb}^{(0)}-{1\over N_b}\rho_{bab}^{(0)}\;{\bf tr}\{MM^+\}=0\;.\end{equation}
Subtracting these equalities from the diagonals of (\ref{TRIP9})
yields the following representative of $\pi_D(G)\in\Omega_D^2({\cal A})$ :
\begin{equation}\label{TRIP15}
\pi_D(G)=\left(\begin{array}{cc}
\pi_D(G)_{[aa]} & \pi_D(G)_{[ab]}\\
\pi_D(G)_{[ba]} & \pi_D(G)_{[bb]}\end{array}\right)\;,\end{equation}
where
\begin{eqnarray*}
\pi_D(G)_{[aa]}&=&-{\bf c}(\rho_{aaa}^{(2)})\otimes{\bf 1}_a
+\rho_{aba}^{(0)}\otimes\Bigl[M^+M\Bigr]_{NT} \;,\\
\pi_D(G)_{[ab]}&=&-{\bf i}\;{\bf c}(\rho_{ab}^{(1)})\gamma_3\otimes M^+\;,\\
\pi_D(G)_{[ba]}&=&-{\bf i}\;{\bf c}(\rho_{ba}^{(1)})\gamma_3\otimes M\;,\\
\pi_D(G)_{[bb]}&=&-{\bf c}(\rho_{bbb}^{(2)})\otimes{\bf 1}_b
+\rho_{bab}^{(0)}\otimes\Bigl[MM^+\Bigr]_{NT}\;.\end{eqnarray*}
with the traceless matrices\footnote{Note that when $N_a=N_b=N$ and $M$ is a scalar matrix,
these traceless matrices $\Bigl[M^+M\Bigr]_{NT}$ and $\Bigl[MM^+\Bigr]_{NT}$ vanish and there is no
$\rho_\alpha^{(0)}$ term in $\pi_D(G)$. Physically this implies that, in order to have a Higgs
mechanism, a nontrivial mass spectrum is necessary!} :
\begin{eqnarray*}
\Bigl[M^+M\Bigr]_{NT}&=&M^+M-{1\over N_a}{\bf tr}\{M^+M\}\;,\\
\Bigl[MM^+\Bigr]_{NT}&=&MM^+-{1\over N_b}{\bf tr}\{MM^+\}\;.\end{eqnarray*}
The scalar product (\ref{TRIP11}) in $\Omega_D^2({\cal A})$ is calculated using,
besides the trace theorems, the identities:
\begin{eqnarray*}{1\over 2^{d/2}}{\bf tr}_{{\bf c}} \Bigl\{({\bf c}(\rho^{(k)}))^+{\bf c}(\rho^{\prime(k)}\Bigr\}
&=&k!\;(\rho^{(k)}_{i_1\dots i_k})^*\rho^{\prime(k)}_{j_1\dots j_k}\;
\delta^{i_1j_1}\dots\delta^{i_kj_k}\\
&=&g^{-1}(\rho^{(k)*};\rho^{\prime(k)}).\end{eqnarray*}
In terms of the Hodge dual $\star$, defined by
\[g^{-1}(\rho^{(k)*};\rho^{\prime(k)})\omega
=(\rho^{(k)})^*\wedge\star\rho^{\prime(k)}\;,\]
the scalar product reads :
\begin{eqnarray}\label{TRIP16}
&&\langle \pi_D(G);\pi_D(G^\prime)\rangle_{2,D}=\nonumber\\
&&{1\over 2\pi}\Biggl\{
N_a\int_{S^2}{\rho^{(2)}_{aaa}}^*\wedge\star\rho^{\prime(2)}_{aaa}
+N_b\int_{S^2}{\rho^{(2)}_{bbb}}^*\wedge\star\rho^{\prime(2)}_{bbb}\nonumber\\
&&
+{\bf tr}\{MM^+\}\;
\int_{S^2}{\rho^{(1)}_{ab}}^*\wedge\star\rho^{\prime(1)}_{ab}
+{\bf tr}\{M^+M\}\;
\int_{S^2}{\rho^{(1)}_{ba}}^*\wedge\star\rho^{\prime(1)}_{ba}\nonumber\\
&&
+{\bf tr}\Bigl\{[M^+M]_{NT}^2\Bigr\}\;
\int_{S^2}{\rho^{(0)}_{aba}}^*\wedge\star\rho^{\prime(0)}_{aba}
+{\bf tr}\Bigl\{[MM^+]_{NT}^2\Bigr\}\;
\int_{S^2}{\rho^{(0)}_{bab}}^*\wedge\star\rho^{\prime(0)}_{bab}\Biggr\}\;.\nonumber\\
&&
\end{eqnarray}
\subsection{The Yang-Mills-Higgs action}\label{Yang-Mills-Higgs}
The universal connection in ${\cal M}$, given by the matrices $(\!( A_{\alpha\beta}(x,y))\!)$
of (\ref{PMod6}), is represented in $\Omega^{(1)}_D({\cal A})$ by an operator of the form (\ref{TRIP8})
where the differential forms $\sigma^{(k)}_{\cdots}$ are matrix-valued.
\begin{eqnarray}\label{YMH1}
\Bigl(\!\Bigl(\sigma_a^{(1)}(x)\Bigr)\!\Bigr)&=&
\Bigl(\!\Bigl(\alpha_a(x)\Bigr)\!\Bigr)\equiv
\Bigl(\vec e_{k,y}(\!(\omega_{aa}(x,y))\!)\Bigr)_{\mid y=x}\;\theta_x^k\;,\nonumber\\
\Bigl(\!\Bigl(\sigma_b^{(1)}(x)\Bigr)\!\Bigr)&=&
\alpha_b(x)(\!( P_b(x))\!)\equiv
\Bigl(\vec e_{k,y}\omega_b(x,y)\Bigr)_{\mid y=x}\;\theta_x^k\;(\!( P_b(x))\!)\;,\nonumber\\
\Bigl(\!\Bigl(\sigma_{ab}^{(0)}(x)\Bigr)\!\Bigr)&=&
\Bigl(\!\Bigl(\vert\Phi_{ab}(x,x)\rangle\langle\nu(x)\vert\Bigr)\!\Bigr)\;,\nonumber\\
\Bigl(\!\Bigl(\sigma_{ba}^{(0)}(x)\Bigr)\!\Bigr)&=&
\Bigl(\!\Bigl(\vert\nu(x)\rangle\langle\Phi_{ba}(x,x)\vert\Bigr)\!\Bigr)\;.
\end{eqnarray}
The monopole connection (\ref{PMod13}) also implements a differential one-form :
\begin{eqnarray}\label{YMH2}
\mu_b(x)&=&(\vec e_{k,y}m_b(x,y))_{\mid y=x}\;\theta^k_x
=\langle\nu(x)\vert\Bigl({\rm d}\vert\nu(x)\rangle\Bigr)\;,\nonumber\\
&=&{1/2\over 1+\vert\nu(x)\vert^2}\Bigl(\nu(x)^*{\rm d}\nu(x)-\nu(x){\rm d}\nu(x)^*\Bigr)\;.
\end{eqnarray}
It is also convenient to introduce the Higgs field doublets :
\begin{eqnarray}\label{YMH3}
\vert\eta_{ab}(x)\rangle&=\vert H_{ab}(x,x)\rangle&
=\vert\Phi_{ab}(x,x)\rangle+\vert\nu(x)\rangle\;,\nonumber\\
\langle\eta_{ba}(x)\vert&=\langle H_{ab}(x,x)\vert&
=\langle\Phi_{ba}(x,x)\vert+\langle\nu(x)\vert\;.\end{eqnarray}
The hermiticity of the connection (\ref{PMod9}) yields :
\begin{equation}\label{YMH4}\Bigl(\!\Bigl(\alpha_a\Bigr)\!\Bigr)^+=-\Bigl(\!\Bigl(\alpha_a\Bigr)\!\Bigr),\;(\alpha_b)^*=-\alpha_b,\;
(\mu_b)^*=-\mu_b,\;\langle\eta_{ba}\vert=\vert\eta_{ab}\rangle^+\;.\end{equation}
From (\ref{PMod17}) it follows that, under an active gauge transformation, the differential forms (\ref{YMH1}) and (\ref{YMH2}) behave as :
\begin{eqnarray*}
(\!(\alpha_a^{\bf U})\!)&=&(\!({\bf U}_a)\!)^{-1}(\!(\alpha_a)\!)(\!({\bf U}_a)\!)+
(\!({\bf U}_a)\!)^{-1}{\rm d}(\!({\bf U}_a)\!)\\
\alpha^{{\bf U}}_b&=&({\bf u}_b)^{-1}\alpha_b({\bf u}_b)+({\bf u}_b)^{-1}{\rm d}{\bf u}_b=
\alpha_b+({\bf u}_b)^{-1}{\rm d}{\bf u}_b\\
\mu^{{\bf U}}_b&=&({\bf u}_b)^{-1}\mu_b({\bf u}_b)=\mu_b\\
\vert\eta^{{\bf U}}_{ab}\rangle
&=&(\!({\bf U}_a)\!)^{-1}\vert\eta_{ab}\rangle {\bf u}_b\;,
\;\langle\eta^{{\bf U}}_{ba}\vert
=({\bf u}_b)^{-1}\langle\eta_{ba}\vert(\!({\bf U}_a)\!)\;.\end{eqnarray*}
On the other hand, under a passive gauge transformation $H_B\rightarrow H_A$, according
to (\ref{PMod10}) and (\ref{PMod14}), they transform as :
\begin{eqnarray*}
&&(\!(\alpha_a^A)\!)=(\!(\alpha_a^B)\!)\;,\;\alpha_b^A=\alpha_b^B\;,\\
&&\vert\eta^A_{ab}\rangle=\vert\eta^B_{ab}\rangle(c_{AB})^{-n/2}\;,\;
\langle\eta^A_{ba}\vert=(c_{AB})^{+n/2}\langle\eta^B_{ba}\vert\;,\\
&&\mu^A_b=\mu^B_b+(c_{AB})^{+n/2}{\rm d}(c_{AB})^{-n/2}=
\mu^B_b-(n/2)(c_{AB})^{-1}{\rm d}c_{AB}\;.\end{eqnarray*}
This means that the Higgs fields \(\{\vert\eta_{ab}\rangle\,;\,\langle\eta_{ba}\vert\}\) are
actually Pensov scalars of weight $\{- n/2\,;\,+n/2\}$ and that the monopole potential cannot be
represented by a globally defined one-form on the sphere, but adquires the inhomogeneous term
\(\;-(n/2)(c_{AB})^{-1}{\rm d}c_{AB}\) in $H_B\cap H_A$.\\
The canonical projection (\ref{TRIP10}) induces a \(\Omega^{(1)}_D({\cal A})\)-valued connection
in ${\cal M}$ :
\[\nabla_D:
{\cal M}\mapsto{\cal M}\otimes_{\cal A}\Omega^{(1)}_D({\cal A}):X\mapsto\nabla_D X\;,\]
defined by \(\nabla_D=\Bigl({\bf 1}_{\cal M}\otimes\pi_D\Bigr)\circ\nabla\).\\
From (\ref{PMod11}) and (\ref{YMH1}) it follows that :
\begin{eqnarray*}
\vert(\nabla_DX)_{aa}\rangle &=&
-{\bf i}\;{\bf c}\biggl({\rm d}\vert f_a\rangle+(\!(\alpha_a)\!) \vert f_a\rangle\biggr)\;,\\
\vert(\nabla_DX)_{ab}\rangle &=&
\biggl(\vert\eta_{ab}\rangle f_b - \vert f_a\rangle\biggr)\gamma_3 M^+\;,\\
\vert(\nabla_DX)_{ba}\rangle &=&\vert \nu\rangle
\biggl(\langle\eta_{ba}\vert f_a\rangle-f_b\biggr)\gamma_3 M \;,\\
\vert(\nabla_DX)_{bb}\rangle &=&\vert\nu\rangle
-{\bf i}\;{\bf c}\biggl({\rm d}f_b +(\alpha_b+\mu_b)f_b\biggr)\;.\end{eqnarray*}
The curvature (\ref{PMod15}) is represented in $\Omega^{(2)}_D({\cal A})$ by $\pi_D(R)$ of the
form (\ref{TRIP15}), where the differential forms $\rho^{(k)}_{\dots}$'s are now $2\times 2$-matrix
valued.\\ The diagonal elements of $\pi_D(R)$ are given by :
\begin{eqnarray*}
\Bigl(\!\Bigl(\rho_{aaa}^{(2)}\Bigr)\!\Bigr)&=
&(\!( F_a)\!)={\rm d}(\!(\alpha_a)\!)+ (\!(\alpha_a)\!)\wedge(\!(\alpha_a)\!)\;,\nonumber\\
\Bigl(\!\Bigl(\rho_{bbb}^{(2)}\Bigr)\!\Bigr)&=
&F_b(\!( P_b)\!)=({\rm d}\alpha_b+{\rm d}\mu_b)(\!( P_b)\!)\;,\nonumber\\
\Bigl(\!\Bigl(\rho_{aba}^{(0)}\Bigr)\!\Bigr)&=
&\vert\eta_{ab}\rangle\langle\eta_{ba}\vert - (\!({\bf Id})\!)\;,\nonumber\\
\Bigl(\!\Bigl(\rho_{bab}^{(0)}\Bigr)\!\Bigr)&=
&(\langle\eta_{ba}\vert\eta_{ab}\rangle - 1)\;(\!( P_b)\!)\;.\end{eqnarray*}
The off-diagonal elements are given in terms of the covariant differentials
of the Higgs fields (\ref{YMH3}):
\begin{eqnarray}\label{YMH5}
\vert\nabla\eta_{ab}\rangle &=&
{\rm d}\;\vert\eta_{ab}\rangle +(\!(\alpha_a)\!)\vert \eta_{ab}\rangle
- \vert \eta_{ab}\rangle(\alpha_b+\mu_b)\;,\nonumber\\
\langle\nabla\eta_{ba}\vert &=&
{\rm d}\;\langle \eta_{ba}\vert -\langle \eta_{ba}\vert (\!(\alpha_a)\!)
+(\alpha_b+\mu_b)\langle \eta_{ba}\vert\;.\end{eqnarray}
They read
\[
\Bigl(\!\Bigl(\rho_{ab}^{(1)}\Bigr)\!\Bigr)=\vert \nabla\eta_{ab}\rangle\langle\nu\vert\;,\;
\Bigl(\!\Bigl(\rho_{ba}^{(1)}\Bigr)\!\Bigr)=\vert\nu\rangle\langle \nabla\eta_{ba}\vert\;.\]
The Yang-Mills-Higgs action is constructed as :
\begin{eqnarray}\label{YMH6}
{\bf S}_{YMH}(\nabla_D)&=&
\lambda\,{\bf tr}_{matrix}\Bigl\{\langle\pi_D(R);\pi_D(R)\rangle_{2,D}\Bigr\}\nonumber\\
&=&\lambda\,{\bf tr}_{matrix}
\biggl\{{\bf Tr}_{Dix}\Bigl\{\pi_D(R)^{\dagger}\pi_D(R)\vert{\bf D}\vert^{-2}\Bigr\}\biggr\}
\;,\end{eqnarray}
where $\lambda$ is a coupling constant and ${\bf tr}_{matrix}$ is the
trace of the $2\times 2$ matrices, product of matrices $\Bigl(\!\Bigl(\rho^{(k)}\Bigr)\!\Bigr)^+$ with $\Bigl(\!\Bigl(\rho^{(k)}\Bigr)\!\Bigr)$. Since the curvature transforms as $ R\rightarrow R^{\bf U}={\bf U}^{-1}R{\bf U}$, the gauge invariance of the action follows from the obvious extension of the representation (\ref{TRIP12}) of the unitaries in ${\cal H}^{(2)}_D$.
With the scalar product given by (\ref{TRIP16}), the action (\ref{YMH5}) reads :
\begin{eqnarray}\label{YMH7}
{\bf S}_{YMH}(\nabla_D)&=&{\lambda\over 2\pi}\biggl\{
N_a\int_{S^2}{\bf tr}_{matrix}\Bigl\{(\!( F_a)\!)^+\wedge\star(\!( F_a)\!)\Bigr\}\nonumber\\&&
\quad\quad+N_b\int_{S^2}(F_b)^*\wedge\star F_b\nonumber\\&&
\quad\quad+2\;{\bf tr}\{MM^+\}\;
\int_{S^2}\langle\nabla\eta_{ba}\vert\wedge\star\vert\nabla\eta_{ab}\rangle\nonumber\\&&
\quad\quad+{\bf tr}\Bigl\{{\Bigl[M^+M\Bigr]_{NT}}^2\Bigr\}\;
\int_{S^2}\star\Bigl((\langle \eta_{ba}\vert \eta_{ab}\rangle-1)^2+1\Bigr)\nonumber\\&&
\quad\quad+{\bf tr}\Bigl\{{\Bigl[MM^+\Bigr]_{NT}}^2\Bigr\}\;
\int_{S^2}\star(\langle \eta_{ba}\vert \eta_{ab}\rangle-1)^2
\biggr\}\;.\end{eqnarray}
\subsection{The Hilbert space of particle states and the covariant Dirac operator}\label{Matter}
The tensor product over ${\cal A}$ of the right ${\cal A}$-module ${\cal M}$ with the (left-module) Hilbert space
${\cal H}$ is itself a Hilbert space ${\cal H}_{\bf p}={\cal M}\otimes_{\cal A}{\cal H}$, with scalar product induced by the scalar product
(\ref{TRIP4}) in ${\cal H}$ and the hermitian structure ${\bf h}$ in the module ${\cal M}$ :
\[(X\otimes_{\cal A} {\bf \Psi};Y\otimes_{\cal A}{\bf \Phi}) = ({\bf \Psi};\pi({\bf h}(X,Y)){\bf \Phi})\;.\]
A generic element of ${\cal H}_{\bf p}$ can be written as \(\Vert{\bf \Psi}_{\bf p}\rangle\!\rangle=E_i\otimes_{\cal A}{\bf \Psi}^i\), where ${\bf \Psi}^i\in{\cal H}$ obeys $\pi({P^i}_j){\bf \Psi}^j={\bf \Psi}^i$.
In the model considered here, ${\cal H}={\cal H}_{(s)}\otimes\Bigl({\bf C}^{N_a}\oplus{\bf C}^{N_b}\Bigr)$ and the projective module is ${\cal M}={\bf P} {\cal A}^2$, with ${\bf P}$ defined by the homotopy class $[g]$ in (\ref{PMod3}). A state \(\Vert{\bf \Psi}_{\bf p}\rangle\!\rangle\) describing particles, is thus represented by :
\begin{enumerate}
\item A pair of Pensov spinors of ${\cal H}_{(s)}$, given by :\\
\(\vert\psi_a(x)\rangle=\left(\begin{array}{c}\psi^{1}_a(x)\\ \psi^{2}_a(x)\end{array}\right)\), each with $N_a$ values of the generation index.
\item A single Pensov spinor $\psi_b(x)$ of ${\cal H}_{(s+n/2)}$, with a $N_b$-valued generation index, such that
\(\vert\psi_b(x)\rangle=\left(\begin{array}{c}\psi^{1}_b(x)\\ \psi^{2}_b(x)\end{array}\right)=\vert\nu(x)\rangle\;\psi_b(x)$ in $H_B$.
\end{enumerate}
The $\star$-representation $\pi$ of $\Omega^\bullet({\cal A})$ in ${\cal H}$ induces a mapping
\begin{equation}\label{Pi1}
\pi_1:{\cal M}\otimes_{\cal A}\Omega^\bullet({\cal A})\mapsto {\cal B}({\cal H},{\cal H}_{\bf p}):
X\otimes_{\cal A} F\mapsto\pi_1(X\otimes_{\cal A} F)\;,\end{equation}
where \({\cal B}({\cal H},{\cal H}_{\bf p})\) are the bounded linear operators from ${\cal H}$ to ${\cal H}_{\bf p}$.\\
It is defined by \(\pi_1(X\otimes_{\cal A} F){\bf \Psi}=X\otimes_{\cal A}\pi(F){\bf \Psi}\;.\)\\
Furthermore, there is a mapping
\begin{equation}\label{Pi2}
\pi_2:{\rm HOM}_{\cal A}({\cal M},{\cal M}\otimes_{\cal A}\Omega^\bullet)\mapsto{\cal B}({\cal H}_{\bf p}):{\bf T}\mapsto\pi_2({\bf T})\;,\end{equation}
defined by \(\pi_2({\bf T})\Bigl(X\otimes_{\cal A}{\bf \Psi}\Bigr)=\pi_1({\bf T}X){\bf \Psi}\).\\
The covariant Dirac operator in ${\cal H}_{\bf p}$ is defined, using (\ref{Pi1}), as
\begin{equation}\label{CovDirac1}
{\bf D}_\nabla\Bigl(X\otimes_{\cal A}{\bf \Psi}\Bigr)=X\otimes_{\cal A}{\bf D}{\bf \Psi}+\pi_1(\nabla X){\bf \Psi}\;.\end{equation}
It is easy to check that \({\bf D}_\nabla\Bigl(X\,f\otimes_{\cal A}{\bf \Psi}\Bigr)=
{\bf D}_\nabla\Bigl(X\otimes_{\cal A}\pi(f){\bf \Psi}\Bigr)\) so that ${\bf D}_\nabla$ is well defined in ${\cal H}_{\bf p}$.\\
A grading in ${\cal H}_{\bf p}$ is defined by
\begin{equation}\label{partgrad}
\Gamma_{\bf p}: X\otimes_{\cal A}{\bf \Psi}\mapsto X\otimes_{\cal A} \Gamma{\bf \Psi}\;\end{equation}
and the covariant Dirac operator is odd with respect to this grading :
\begin{equation}\label{oddness}
{\bf D}_\nabla\Gamma_{\bf p}+\Gamma_{\bf p}{\bf D}_\nabla=0
\end{equation}
With $\Vert{\bf \Psi}_{\bf p}\rangle\!\rangle$ as above, \({\bf D}_\nabla\) is calculated as follows.
\[{\bf D}_\nabla\Vert{\bf \Psi}_{\bf p}\rangle\!\rangle=
E_i\otimes_{\cal A}\Bigl({\cal D}_\nabla\Vert{\bf \Psi}_{\bf p}\rangle\!\rangle\Bigr)^i\;,\]
where
\(\Bigl({\cal D}_\nabla\Vert{\bf \Psi}_{\bf p}\rangle\!\rangle\Bigr)^i=
\Bigl(\pi({P^i}_j){\bf D}+\pi({(\!( A)\!)^i}_j)\Bigr){\bf \Psi}^j\) is represented by
\begin{equation}\label{CovDirac2}
\left(\begin{array}{cc}
({\cal D}_\nabla)_{aa}&({\cal D}_\nabla)_{ab}\\
({\cal D}_\nabla)_{ba}&({\cal D}_\nabla)_{bb}\end{array}\right)
\left(\begin{array}{l}\vert\psi_a\rangle\\ \vert\nu\rangle\psi_b\end{array}\right)\;,\end{equation}
\begin{eqnarray*}
\Bigl({\cal D}_\nabla\Bigr)_{aa}&=&
{\cal D}_{(s)}\otimes{\bf 1}_a-{\bf i}\;{\bf c}(\!(\alpha_a)\!)\otimes{\bf 1}_a\;,\\
\Bigl({\cal D}_\nabla\Bigr)_{ab}&=&
(\!(\vert\eta_{ab}\rangle\langle\nu\vert)\!)\otimes\gamma_3 M^+\;,\\
\Bigl({\cal D}_\nabla\Bigr)_{ba}&=&
(\!(\vert\nu\rangle\langle\eta_{ba}\vert)\!)\otimes\gamma_3 M\;,\\
\Bigl({\cal D}_\nabla\Bigr)_{bb}&=&
(\!( P_b)\!){\cal D}_{(s)}(\!( P_b)\!)
\otimes{\bf 1}_b-{\bf i}\;{\bf c}(\alpha_b)(\!( P_b)\!)\;.\end{eqnarray*}
Now, \(\langle\nu\vert{\cal D}_{(s)}\vert\nu\rangle\psi_b={\cal D}_{(s)}\psi_b-{\bf i}\;{\bf c}(m_b)\psi_b\) and
with our choice (\ref{PMod3}) of the representative of the homotopy class $[n]\in{\bf Z}$,
we obtain\footnote{Note that with a different choice in (\ref{PMod3}), a globally defined
differential one-form would be added to \({\cal D}_{(s+n/2)}\) and this can always be absorbed in
\(\alpha_b\).}
\begin{equation}\label{SfromI}
(\!( P_b)\!){\cal D}_{(s)}\vert\nu\rangle\psi_b=\vert\nu\rangle{\cal D}_{(s+n/2)}\psi_b\;.\end{equation}
Substituting (\ref{SfromI}) and (\ref{CovDirac2}) in (\ref{CovDirac1}) yields finally:
\begin{eqnarray}\label{CovDirac3}
\Bigl({\cal D}_\nabla\Vert{\bf \Psi}_{\bf p}\rangle\!\rangle\Bigr)_a&=&
\Bigl({\cal D}_{(s)}-{\bf i}\;{\bf c}(\!(\alpha_a)\!)\Bigr)\vert\psi_a\rangle
+\vert \eta_{ab}\rangle\gamma_3 M^+\psi_b\;,\nonumber\\
\Bigl({\cal D}_\nabla\Vert{\bf \Psi}_{\bf p}\rangle\!\rangle\Bigr)_b&=&
\vert\nu\rangle\Bigl[\Bigl({\cal D}_{(s+n/2)}
-{\bf i}\;{\bf c}(\alpha_b)\Bigr)\psi_b
+\gamma_3 M\langle \eta_{ba}\vert\psi_a\rangle\Bigr]\;.\end{eqnarray}
The matter action functional is then constructed as ;
\begin{eqnarray}\label{Matiere}
{\bf S}_{\bf Mat}(\Vert{\bf \Psi}_{\bf p}\rangle\!\rangle,\nabla_D)&=&
\Bigl(\Vert{\bf \Psi}_{\bf p}\rangle\!\rangle;{\bf D}_\nabla\Vert{\bf \Psi}_{\bf p}\rangle\!\rangle\Bigr)\nonumber\\
&=&\int_{S^2}\star\biggl\{
\langle\psi_a\vert\Bigl({\cal D}_{(s)}-{\bf i}\;{\bf c}(\!(\alpha_a)\!)\Bigr)
\vert\psi_a\rangle\nonumber\\
&&+\langle\psi_a\vert\eta_{ab}\rangle\gamma_3 M^+\psi_b
+(\psi_b)^+\gamma_3 M\langle \eta_{ba}\vert\psi_a\rangle\nonumber\\
&&+(\psi_b)^+\Bigl({\cal D}_{(s+n/2)}-{\bf i}\;{\bf c}(\alpha_b)\Bigr)\psi_b
\biggr\}\;.\end{eqnarray}
The hermiticity condition (\ref{YMH4}) guarantees that the action is real.
If an Euclidean chiral theory is aimed for, then
\(\Gamma_{\bf p}\Vert{\bf \Psi}_{\bf p}\rangle\!\rangle=\Vert{\bf \Psi}_{\bf p}\rangle\!\rangle\) implies that the action
(\ref{Matiere}) vanishes identically due to the oddness of the Dirac operator (\ref{oddness}).
A proposed way out, as in \cite{V-GB}, is just to make an easy switch going to an indefinite
Minkowski type metric changing the $\psi^+$ to a $\bar\psi=\psi^+ \gamma^0$ so that the presence
of the $\gamma^0$ provides an extra factor minus one and the action does not vanish. Such a "usual
incantation " \cite{M-GB-V} appears highly unaesthetic and rather unsatisfactory. It seems
necessary to double the Hilbert space in order to deal with this issue. This can be achieved
introducing a Hilbert space ${\cal H}_{\bf \bar p}$ of "anti-particle" states. The need of doubling the
Fermion fields also arises in the usual Euclidean quantum field theory,
where the fermion fields are operator valued in Fock space, in order to cure inconsistent
hermiticity properties of the propagators \cite{O-S,F-O}. Alternative proposals were made by
\cite{Met} and more recently by \cite{vN-W}. Related comments
by \cite{L-M-M-S} in a non-commutative geometric setting, should also be mentioned.\\
Here, however, we choose to remain with the primary interpretation of \(\Vert{\bf \Psi}_{\bf p}\rangle\!\rangle\)
as a {\it state} in the Hilbert space \({\cal H}_{\bf p}\) represented by Euclidean {\it wave
functions}. This means that in this work we endeavour an Euclidean one-particle (plus one would-be
anti-particle) field theory, which, in a path integral formalism, may hopefully lead to a proper
quantum theory.
\newpage
\section{Real spectral triples}\label{Real Triples}
\setcounter{equation}{0}
\subsection{The real Pensov-Dirac spectral triple}\label{RPDtriple}
The complex conjugation ${\cal K}$ transforms a Pensov field of weight $s$, $\sigma^{(s)}$, into a
Pensov field $\sigma^{(s)*}$ of weight $-s$. Besides the Hilbert space of Pensov spinors
\(\psi_{(+)}\) of weight $s$, denoted here as
${\cal H}_{1(+)}$, we also introduce ${\cal H}_{1(-)}$, with spinors \(\psi_{(-)}\) of weight $-s$.
A real structure will be induced by a pair of anti-linear mappings
${\bf J}_{1(\pm)}:{\cal H}_{1(\pm)}\mapsto{\cal H}_{1(\mp)}$, which are required to preserve the real
Clifford-algebra module structure of the spinor spaces :
\begin{eqnarray}
{\bf J}_{1(\pm)}\lambda\psi_{(\pm)}&=&\bar\lambda{\bf J}_{1(\pm)}\psi_{(\pm)}\nonumber\\
{\bf J}_{1(\pm)}\left(\gamma^k\psi_{(\pm)}\right)&=
&\alpha\;\gamma^k\left({\bf J}_{1(\pm)}\psi_{(\pm)}\right)\label{J1}\;,\end{eqnarray}
where we allow for $\alpha$ to be a sign factor $\pm 1$.\\
With
\begin{equation}\label{J2}
{\bf J}_{1(\pm)}\psi_{(\pm)}=a_{(\pm s)}\;{\cal C}_{1,\alpha}\,{\cal K}\;\psi_{(\pm)}\;,\end{equation}
where $a_{(\pm s)}$ is an arbitrary complex number, we should have
\[{\cal C}_{1,\alpha}\left(\gamma^k\right)^*{\cal C}_{1,\alpha}^{-1}=\alpha\;\gamma^k\;.\]
In the chiral representation
\(\gamma^1=\left(\begin{array}{cc} 0 & 1\\ 1 &
0\end{array}\right)\;;\;\gamma^2=\left(\begin{array}{cc} 0 & {\bf i}\\ -{\bf i} &
0\end{array}\right)\), we may choose
${\cal C}_{1,\alpha}=\left(\begin{array}{cc} 0 & 1\\ \alpha & 0
\end{array}\right)$. \\ The adjoint of \({\bf J}_{1(\pm)}\) is given by :
\begin{equation}\label{J3}
\left({\bf J}_{1(\pm)}\right)^+\psi_{(\mp)}=a_{(\pm s)}\;{\cal C}_{1,\alpha}^t\,{\cal K}\;\psi_{(\mp)}\;.\end{equation}
Demanding that
\begin{equation}\label{J4}
{\bf J}_{1(\pm)}\left({\bf J}_{1(\pm)}\right)^+=
{\bf 1}_{(\mp)}\;,\;\left({\bf J}_{1(\pm)}\right)^+{\bf J}_{1(\pm)}={\bf 1}_{(\pm)}\;,\end{equation}
restricts $a_{(\pm s)}$ be a phase factor. On \({\cal H}_{1}={\cal H}_{1(+)}\oplus{\cal H}_{1(-)}\), the antilinear isometry defined by
\begin{equation}\label{J5}
{\bf J}_{1}=\left(\begin{array}{cc} 0 &{\bf J}_{1(-)} \\ {\bf J}_{1(+)} & 0 \end{array}\right)\;\end{equation}
obeys
\[{\bf J}_{1}\left(\begin{array}{cc}
\gamma^k & 0 \\ 0 & \gamma^k\end{array}\right)= \alpha \;
\left(\begin{array}{cc}
\gamma^k & 0 \\ 0 & \gamma^k\end{array}\right){\bf J}_{1}\;.\]
If we require
\begin{equation}\label{J6}
{{\bf J}_{1}}^2=\epsilon_1\;{\bf 1}_1\;,\;\hbox{with}\;\epsilon_1=\pm 1,\end{equation}
the phases $a_{(\pm s)}$ are related by $a_{(-s)}=\alpha\;\epsilon_1\;a_{(+s)}$ and
\({\bf J}_{1(-)}=\epsilon_1\left({\bf J}_{1(+)}\right)^+\).
The antilinear mappings ${\bf J}_{1(\pm)}$ intertwine with the Dirac operators ${\cal D}_{(\pm s)}$ as :
\begin{equation}\label{J7}
{\bf J}_{1(\pm)}{\cal D}_{(\pm s)}=-\alpha\;{\cal D}_{(\mp s)}{\bf J}_{1(\pm)}\;.
\end{equation}
On ${\cal H}_{1(+)}$, the Dirac operator is chosen as ${\cal D}_{1(+)}={\cal D}_{(s)}$, but on ${\cal H}_{1(-)}$ we may choose ${\cal D}_{(-s)}$ up to a sign. Let $\epsilon_1^{\;\prime}$ be another arbitrary sign factor, then the choice\footnote{For notational convenience, we define $\kappa(+)=+1$ and $\kappa(-)=-\alpha\;\epsilon_1^{\;\prime}$ so that ${\cal D}_{1(\pm)}=\kappa(\pm)\;{\cal D}_{(\pm s)}$.}
\begin{equation}\label{J8}
{\cal D}_{1(-)}=-\alpha\;\epsilon_1^{\;\prime}\;{\cal D}_{(-s)}\;,\end{equation}
yields a Dirac operator
\({\cal D}_1=\left(\begin{array}{cc}{\cal D}_{1(+)} & 0 \\ 0 & {\cal D}_{1(-)}\end{array}\right)\)
intertwining with ${\bf J}_1$ as :
\begin{equation}\label{J9}
{\bf J}_1\;{\cal D}_1=\epsilon_1^{\;\prime}\;{\cal D}_1{\bf J}_1\;.\end{equation}
The representation of ${\cal A}_1={\cal C}(S^2;{\bf C})$ in ${\cal H}_{1}$ is obtained by taking two copies of the representation in ${\cal H}_{(s)}$ :
\begin{equation}\label{J10}
\pi_1(f)\;\left(\begin{array}{c}\psi_{(+)}\\ \psi_{(-)}\end{array}\right)(x)=
\left(\begin{array}{cc} f(x) & 0 \\
0 & f(x) \end{array}\right)
\left(\begin{array}{c}\psi_{(+)}(x)\\ \psi_{(-)}(x)\end{array}\right)\;.\end{equation}
In general, a real structure ${\bf J}_{1}$ induces a representation of the opposite algebra ${\cal A}_1^{o}$ by :
\[\pi^{o}_{1}(f)={\bf J}_{1}\left(\pi_1(f)\right)^+{\bf J}_{1}^+\;,\]
so that the Hilbert space ${\cal H}_{1}$ becomes an ${\cal A}_1$ bimodule. Here ${\cal A}_1$ is abelian, and with the representation $\pi_1$ above, we have $\pi^{o}_1(f)=\pi_1(f)$.\\
Since
\(\Bigl[D_{1},\pi_{1}(f)\Bigl]=
\left(\begin{array}{cc} -{\bf i}\kappa_{(+)}{\bf c}(df) & 0 \\ 0 & -{\bf i}\kappa_{(-)}{\bf c}(df)\end{array}\right)\), the first-order condition
\begin{equation}\label{J11}\Bigl[\Bigl[{\cal D}_{1},\pi_{1}(f)\Bigr],\pi_{1}^{o}(g)\Bigr]=0\;,\end{equation}
which is needed to define a connection in bimodules \cite{Dubois}, is satisfied.\\
Since \({\bf J}_{1(\pm)}\gamma_3=-\gamma_3{\bf J}_{1(\pm)}\), the chirality in ${\cal H}_{1(+)}\oplus{\cal H}_{1(-)}$ will be taken as
\begin{equation}\label{J12}
\chi_{1}=\left(\begin{array}{cc}\gamma_3 & 0 \\ 0 & \gamma_3 \end{array}\right)\;.\end{equation}
With this choice :
\begin{equation}\label{J13}
{\bf J}_{1}\chi_{1}=\epsilon_1^{\;\prime\prime}\;\chi_{1}{\bf J}_{1}\;,\;\hbox{with}\;\epsilon_1^{\;\prime\prime}=-1\;.\end{equation}
The $\epsilon$-sign table of Connes \cite{Con2,Con3}, corresponding to $n=2$, can be satisfied if we choose
$\epsilon_1=-1$ and $\epsilon_1^{\;\prime}=+1$, but for the moment we shall leave these choices open.\\
The spectral triple \({\cal T}_1=\left\{{\cal A}_1,{\cal H}_{1},{\cal D}_{1},\chi_{1},{\bf J}_1\right\}\) is actually a 0-sphere real spectral triple as defined in \cite{Con2}. For our pragmatic purposes, an $S^0$-real spectral triple may be defined as a real spectral triple with an hermitian involution $\sigma^0$ commuting with $\pi({\cal A}_1),{\cal D}_{1},\chi_{1}$ and anticommuting with ${\bf J}_{1}$.\\
It is implemented by the decomposition ${\cal H}_{1}={\cal H}_{1(+)}\oplus{\cal H}_{1(-)}$, which it is given in, by :
\begin{equation}\label{J14}
\sigma^0=\left(\begin{array}{cc}
{\bf 1}_{1(+)} & 0\\
0 & -{\bf 1}_{1(-)}
\end{array}\right)\;.\end{equation}
The doubling of the Hilbert space is justified if we interpret the Pensov spinors of ${\cal H}_{(s)}$ as usual (Euclidean!) Dirac spinors interacting with a magnetic monopole of strenght $s$. It seems then natural to consider the (Euclidean!) anti-particle fields as Dirac spinors "seeing" a monopole of strenght $-s$ i.e. as Pensov spinors of ${\cal H}_{(-s)}$.
\subsection{The real discrete spectral triple}\label{Rdtriple}
Proceeding further, as in section {\bf\ref{Triple}}, we have to compose the above $S^0$-real "Dirac-Pensov" spectral triple
\({\cal T}_1\) with a real discrete spectral triple \({\cal T}_2=\left\{{\cal A}_2,{\cal H}_{2},{\cal D}_{2},\chi_{2},{\bf J}_2\right\}\) over the algebra \({\cal A}_2={\bf C}\oplus{\bf C}\).
The most general finite Hilbert space allowing a \({\cal A}_2\)-bimodule structure\footnote{For a general discussion on real discrete spectral triples, we refer to (\cite{Kraj},\cite{Pasch})} is given by the direct sum
\begin{equation}\label{J15}
{\cal H}_2=\bigoplus_{\alpha,\beta} {\bf C}^{\;N_{\alpha\beta}}\;,\end{equation}
where $\alpha$ and $\beta$ vary over \{a,b\} and where $N_{\alpha\beta}$ are integers.\\
Its elements are of the form
\[\left(\xi\right)=\left(\begin{array}{c}\xi^{aa} \\ \xi^{ab} \\ \xi^{ba} \\ \xi^{bb}\end{array}\right)\;,\]
where each $\xi^{\alpha\beta}$ is a column vector with $N_{\alpha\beta}$ rows.
An element $\lambda=(\lambda_a,\lambda_b)$ of ${\cal A}_{1}$ acts on the left on ${\cal H}_{2}$ by:
\begin{equation}\label{J16}
\pi_2(\lambda)=
\left(\begin{array}{cccc}
\lambda_a\;{\bf 1}_{N_{aa}} & 0 & 0 & 0\\
0 & \lambda_a\;{\bf 1}_{N_{ab}} & 0 & 0\\
0 & 0 &\lambda_b\;{\bf 1}_{N_{ba}} & 0\\
0 & 0 & 0 & \lambda_b\;{\bf 1}_{N_{bb}}
\end{array}\right)\;,\end{equation}
and on the right by:
\begin{equation}\label{J17}
\pi_2^{o}(\lambda)=
\left(\begin{array}{cccc}
\lambda_a\;{\bf 1}_{N_{aa}} & 0 & 0 & 0\\
0 & \lambda_b\;{\bf 1}_{N_{ab}} & 0 & 0\\
0 & 0 &\lambda_a\;{\bf 1}_{N_{ba}} & 0\\
0 & 0 & 0 & \lambda_b\;{\bf 1}_{N_{bb}}
\end{array}\right)\;.\end{equation}
Although ${\cal A}_{2}$ is an abelian algebra and, as such, isomorphic to its opposite algebra , it is not a simple algebra so that, in general, $\pi_2(\lambda)\neq\pi_2^{o}(\lambda)$. The discrete real structure, given by ${\bf J}_{2}={\cal C}_{2}{\cal K}$, relates both by
$\pi^{o}(\lambda)={\bf J}_{2}\pi(\lambda)^+{\bf J}_{2}^{-1}$ so that ${\cal C}_{2}$ is an intertwining operator for the two representations:
\begin{equation}\label{J18}
\pi_2^{o}(\lambda)={\cal C}_{2}\pi_2(\lambda){\cal C}_{2}^{-1}\;.\end{equation}
This requires that $N_{ab}=N_{ba}\doteq N$ and, since we require ${\bf J}_2$ to be anti-unitary, the basis in ${\cal H}_2$ may be chosen such that :
\begin{equation}\label{J19}
{\cal C}_{2}=\left(\begin{array}{cccc}
{\bf 1}_{N_{aa}} & 0 & 0 & 0 \\
0 & 0 & {\bf 1}_{N} & 0 \\
0 & {\bf 1}_{N} & 0 & 0 \\
0 & 0 & 0 & {\bf 1}_{N_{bb}}
\end{array}\right)\;.\end{equation}
This implies that :
\begin{equation}\label{J20}
{{\bf J}_2}^2=\epsilon_2\;{\bf 1}_2\;,\;\hbox{with}\;\epsilon_2=+1\;.\end{equation}
The chirality, $\chi_{2}$, defining the orientation of the spectral triple is the image of a Hochschild 0-cycle, i.e. an element of ${\cal A}_{2}\otimes{\cal A}^{o}_{2}$. This implies that $\chi_{2}$ is diagonal and $\chi_{\alpha\beta}=\pm 1$ on each subspace ${\bf C}^{N_{\alpha\beta}}$.\\
Furthermore, demanding that\footnote{If we should require that $\epsilon_2^{\;\prime\prime}=-1$, then $N_{aa}=N_{bb}=0$ and $\chi_{ab}=-\chi_{ba}$ and the corresponding odd Dirac operator would not satisfy the first order condition.}
\begin{equation}\label{J21}
{\bf J}_{2}\chi_{2}=\epsilon_2^{\;\prime\prime}\chi_{2}{\bf J}_{2}\;,\;\hbox{with}\;\epsilon_2^{\;\prime\prime}=+1\;,\end{equation}
requires $\chi_{ab}=\chi_{ba}=\chi^\prime$ so that the chirality in ${\cal H}_{2}$ reads:
\begin{equation}\label{J22}
\chi_{2}=\left(\begin{array}{cccc}
\chi_{aa}{\bf 1}_{N_{aa}} & 0 & 0 & 0\\
0 & \chi^\prime{\bf 1}_{N} & 0 & 0\\
0 & 0 & \chi^\prime{\bf 1}_{N} & 0\\
0 & 0 & 0 & \chi_{bb}{\bf 1}_{N_{bb}}
\end{array}\right)\;.\end{equation}
We consider the following three possibilities leading to a non trivial hermitian Dirac operator, odd with respect to this chirality :
\begin{description}
\item [2.a] $+\chi_{aa}=+\chi^\prime=-\chi_{bb}=\pm 1$
\item [2.b] $-\chi_{aa}=+\chi^\prime=+\chi_{bb}=\pm 1$
\item [2.c] $-\chi_{aa}=+\chi^\prime=-\chi_{bb}=\pm 1$
\end{description}
The corresponding Dirac operators have the form
\begin{description}
\item [2.a,2.b]
\[{\cal D}_{2.a}=\left(\begin{array}{cccc}
0 & 0 & 0 & K^+\\
0 & 0 & 0 & A^+\\
0 & 0 & 0 & B^+\\
K & A & B & 0
\end{array}\right)\;;\;
{\cal D}_{2.b}=\left(\begin{array}{cccc}
0 & B^+ & A^+ & K^+\\
B & 0 & 0 & 0\\
A & 0 & 0 & 0\\
K & 0 & 0 & 0
\end{array}\right)\]
\item [2.c]
\[{\cal D}_{2.c}=\left(\begin{array}{cccc}
0 & B^{\prime +} & A^{\prime +} & 0\\
B^\prime & 0 & 0 & A^+\\
A^\prime & 0 & 0 & B^+\\
0 & A & B & 0
\end{array}\right)\;.\]
\end{description}
The first-order condition $[[{\cal D}_{2},\pi_2(\lambda)],\pi_2^{o}(\mu)]=0$, satisfied in case {\bf 2.c)}, implies that in case {\bf 2.a)} and {\bf 2.b)} $K$ must vanish.\\
If we asssume
\begin{equation}\label{J23}{\bf J}_{2}{\cal D}_{2}=\epsilon_2^\prime\;{\cal D}_{2}{\bf J}_{2}\;,
\;\hbox{with}\;\epsilon_2^\prime=+1\;,\end{equation}
then
\begin{equation}\label{J24}
B=A^*\;;\,B^\prime=A^{\prime\,*}\;.\end{equation}
It should be stressed that, in order to have a non trivial Dirac operator, necessarily $N\neq 0$. This confirms that the discrete Hilbert space ${\cal H}_{dis}$ used in {\bf \ref{Triple}} does not allow for a real structure in the above sense. At last, it can be shown\cite{Kraj,Pasch} that noncommutative Poincar\'{e} duality, in the discrete case, amounts to the non degeneracy of the intersection matrix with elements $\bigcap_{\alpha\beta}=\chi_{\alpha\beta}N_{\alpha\beta}$. This non degeneracy condition in case {\bf 2.a)} and {\bf 2.b)} reads $N_{aa}N_{bb}+N^2\neq 0$ and is always satisfied.
In case {\bf 2.c)} it is required that $N_{aa}N_{bb}-N^{2}\neq 0$ and if all $N$'s should be equal, this would not be satisfied. If we insist on equal $N$'s, which is not strictly necessary, we are limited to the models {\bf 2.a)} and {\bf 2.b)} with representations given by (\ref{J16}), (\ref{J17}), Dirac operators and chirality assignments by :
\begin{equation}\label{J25}
{\cal D}_{2.a}=\left(\begin{array}{cccc}
0 & 0 & 0 & 0\\
0 & 0 & 0 & A^+\\
0 & 0 & 0 & B^+\\
0 & A & B & 0
\end{array}\right)\;;\;
\chi_{2.a}=\chi^{\;\prime}\left(\begin{array}{cccc}
{\bf 1}_{N} & 0 & 0 & 0\\
0 & {\bf 1}_{N} & 0 & 0\\
0 & 0 & {\bf 1}_{N} & 0\\
0 & 0 & 0 & -{\bf 1}_{N}
\end{array}\right)\;.\end{equation}
\begin{equation}\label{J26}
{\cal D}_{2.b}=\left(\begin{array}{cccc}
0 & B^+ & A^+ & 0\\
B & 0 & 0 & 0\\
A & 0 & 0 & 0\\
0 & 0 & 0 & 0
\end{array}\right)\;;\;
\chi_{2.b}=\chi^{\;\prime}\left(\begin{array}{cccc}
-{\bf 1}_{N} & 0 & 0 & 0\\
0 & {\bf 1}_{N} & 0 & 0\\
0 & 0 & {\bf 1}_{N} & 0\\
0 & 0 & 0 & {\bf 1}_{N}
\end{array}\right)\;.\end{equation}
\subsection{The product}\label{Prodtriple}
The product of \({\cal T}_1\) with \({\cal T}_2\) yields the total triple \({\cal T}=\left\{{\cal A},{\cal H},{\cal D},\chi,{\bf J}\right\}\)
with algebra ${\cal A}={\cal A}_1\otimes{\cal A}_2$. Since ${\cal T}_1$ is $S^0$-real, the product is also $S^0$-real with hermitian involution \(\Sigma^0=\sigma^0\otimes{\bf 1}_2\). The total Hilbert space ${\cal H}={\cal H}_1\otimes{\cal H}_2$ is decomposed as
\begin{equation}\label{J27}
{\cal H}={\cal H}_{(+)}\oplus{\cal H}_{(-)}\;,\end{equation}
where ${\cal H}_{(\pm)}={\cal H}_{1(\pm)}\otimes{\cal H}_2$. Elements of ${\cal H}_{(\pm)}$ are represented by column matrices of the form
\begin{equation}\label{J28}
\Psi_{(\pm)}(x)=\left(\begin{array}{c}
\psi_{(\pm)aa}(x) \\ \psi_{(\pm)ab}(x) \\ \psi_{(\pm)ba}(x) \\ \psi_{(\pm)bb}(x) \end{array}\right)\;,\end{equation}
where each $\psi_{(\pm)\alpha\beta}(x)$ is a Pensov spinor of ${\cal H}_{1(\pm)}$ with $N$ "internal" indices (not explicitely written down).\\
The total Dirac operator is
\begin{equation}\label{J29}
{\cal D}={\cal D}_1\otimes{\bf 1}_2+\chi_1\otimes{\cal D}_2\;,\end{equation}
and the chirality is given by
\begin{equation}\label{J30}
\chi=\chi_1\otimes\chi_2\;.\end{equation}
The continuum spectral triple ${\cal T}_1$ of {\bf \ref{RPDtriple}} is of dimension two and its real structure ${\bf J}_1$ obeys
\[{{\bf J}_1}^2=\epsilon_1\;{\bf 1}_1\;;\;{\bf J}_1{\cal D}_1=\epsilon_1^{\;\prime}{\cal D}_1{\bf J}_1\;;\; {\bf J}_1\chi_1=\epsilon_1^{\;\prime\prime}\chi_1{\bf J}_1\;,\]
where $\epsilon_1^{\;\prime\prime}=-1$ was fixed but $\epsilon_1$ and $\epsilon_1^{\;\prime}$ were independent free $\pm 1$ factors. \\
On the other hand the discrete triple ${\cal T}_2$ of {\bf \ref{Rdtriple}} is of zero dimension and ${\bf J}_2$ obeys
\[{{\bf J}_2}^2=\epsilon_2\;{\bf 1}_2\;;\;{\bf J}_2{\cal D}_2=\epsilon_2^{\;\prime}{\cal D}_2{\bf J}_2\;;\; {\bf J}_2\chi_2=\epsilon_2^{\;\prime\prime}\chi_2{\bf J}_2\;,\] with $\epsilon_2=\epsilon_2^{\;\prime}=\epsilon_2^{\;\prime\prime}=+1$.\\
The real structure ${\bf J}$ of the product triple should obey
\begin{equation}\label{J31}
{{\bf J}}^2=\epsilon\;{\bf 1}\;;\;{\bf J}{\cal D}=\epsilon^{\;\prime}{\cal D}{\bf J}\;;\;{\bf J}\chi=\epsilon^{\;\prime\prime}\chi{\bf J}\;.\end{equation}
If we require that Connes' sign table be satisfied, i.e.
\begin{itemize}
\item for ${\cal T}_1$, $n_1=2$ and $\epsilon_1=-1\;,\;\epsilon_1^{\;\prime}=+1\;,\;\epsilon_1^{\;\prime\prime}=-1$,
\item for ${\cal T}_2$, $n_2=0$ and $\epsilon_2=\;\epsilon_2^{\;\prime}=\epsilon_2^{\;\prime\prime}=+1$ for $n_2=0$,
\item for the product ${\cal T}$, $n=2$ and $\epsilon=-1\;,\;\epsilon^{\;\prime}=+1\;,\;\epsilon^{\;\prime\prime}=-1$,
\end{itemize}
it is seen that, with ${\bf J}={\bf J}_1\otimes{\bf J}_2$, the sign table for the product is not obeyed, since such a ${\bf J}$ implies the consistency conditions
\begin{eqnarray}\label{J32}
\epsilon&=&\epsilon_1\;\epsilon_2\;,\nonumber\\
\epsilon^{\;\prime}&=&\epsilon^{\;\prime}_1=\epsilon^{\;\prime\prime}_1\;\epsilon^{\;\prime}_2\;,\nonumber\\
\epsilon^{\;\prime\prime}&=&\epsilon^{\;\prime\prime}_1\;\epsilon^{\;\prime\prime}_2\;,
\end{eqnarray}
and the second condition is not satisfied. If we keep the same Dirac operator (\ref{J29}), it is the definition of ${\bf J}$ that should be changed\footnote{A general examination of the $\epsilon$ sign table and the product of two real spectral triples is done in \cite{Van}.} to
\begin{equation}\label{J33}
{\bf J}={\bf J}_1\otimes({\bf J}_2\;\chi_2)\;,\end{equation}
and with this ${\bf J}$ the consistency conditions become
\begin{eqnarray}\label{J34}
\epsilon&=&\epsilon_1\;\epsilon_2\;\epsilon_2^{\;\prime\prime},\nonumber\\
\epsilon^{\;\prime}&=&\epsilon^{\;\prime}_1=-\epsilon^{\;\prime\prime}_1\;\epsilon^{\;\prime}_2\;,\nonumber\\
\epsilon^{\;\prime\prime}&=&\epsilon^{\;\prime\prime}_1\;\epsilon^{\;\prime\prime}_2\;,
\end{eqnarray}
and these are satisfied. In the rest of this section, we shall assume that these choices are made. Also, in order to simplify the forthcoming formulae, we take $\alpha=-1$ so that
${\cal C}_1=\left(\begin{array}{cc} 0 & 1\\ -1 & 0 \end{array}\right)$ and $a_{(+s)}=a_{(-s)}=+1$ which imply that
\({\bf J}_{1(-)}=-{{\bf J}_{1(+)}}^+={\bf J}_{1(+)}\) and ${\cal D}_{1(\pm)}={\cal D}_{(\pm s)}$. The change of ${\bf J}_2$ to ${\bf J}_2\chi_2$ will not change the representation $\pi_2^{o}$ since $\pi_2$ is even with respect to $\chi_2$.\\
The $S^0$-real structure implies that $\pi$ and $\pi^{o}$, ${\cal D}$ and $\chi$ are block diagonal in the decomposition
(\ref{J27}) of ${\cal H}$ and we obtain
\begin{description}
\item[the representations :] let $f\in{\cal C}(S^2,{\bf C}\oplus{\bf C})$, then
\begin{equation}\label{J35}
\pi(f)=\pi_{(+)}(f)\oplus\pi_{(-)}(f)\;;\;\pi^{o}(f)=\pi^{o}_{(+)}(f)\oplus\pi^{o}_{(-)}(f)\;,\end{equation}
\begin{eqnarray}
\pi_{(\pm)}(f(x))& = &
\left(\begin{array}{cccc}
f_a(x){\bf 1}_N & 0 & 0 & 0 \\
0 & f_a(x){\bf 1}_N & 0 & 0 \\
0 & 0 & f_b(x){\bf 1}_N & 0 \\
0 & 0 & 0 & f_b(x){\bf 1}_N
\end{array}\right)\;,\nonumber\\
&&\nonumber\\
\pi^{o}_{(\pm)}(f(x))& = &
\left(\begin{array}{cccc}
f_a(x){\bf 1}_N & 0 & 0 & 0 \\
0 & f_b(x){\bf 1}_N & 0 & 0 \\
0 & 0 & f_a(x){\bf 1}_N & 0 \\
0 & 0 & 0 & f_b(x){\bf 1}_N
\end{array}\right)\;.\nonumber\\
&&\label{J36}\end{eqnarray}
\item[the Dirac operator :] \({\cal D}={\cal D}_{(+)}\oplus{\cal D}_{(-)}\) is given by
\begin{itemize}
\item in case {\bf 2.a)} :
\begin{equation}\label{J37}
{\cal D}_{{\bf a}(\pm)}=\left(\begin{array}{cccc}
{\cal D}_{1(\pm)}{\bf 1}_N & 0 & 0 & 0 \\
0 & {\cal D}_{1(\pm)}{\bf 1}_N& 0 & \gamma_{3}A^+ \\
0 & 0 & {\cal D}_{1(\pm)}{\bf 1}_N & \gamma_{3}B^+ \\
0 & \gamma_{3}A & \gamma_{3} B & {\cal D}_{1(\pm)}{\bf 1}_N \end{array}\right)\;,\end{equation}
\item in case {\bf 2.b)} :
\begin{equation}\label{J38}
{\cal D}_{{\bf b}(\pm)}=\left(\begin{array}{cccc}
{\cal D}_{1(\pm)}{\bf 1}_N & \gamma_{3}B^+ & \gamma_{3}A^+ & 0 \\
\gamma_{3}B & {\cal D}_{1(\pm)}{\bf 1}_N& 0 & 0 \\
\gamma_{3}A & 0 & {\cal D}_{1(\pm)}{\bf 1}_N & 0 \\
0 & 0 & 0 & {\cal D}_{1(\pm)}{\bf 1}_N \end{array}\right)\;.\end{equation}
\end{itemize}
\item[the chirality :] \(\chi=\chi_{(+)}\oplus\chi_{(-)}\), is given by
\begin{itemize}
\item in case {\bf 2.a)} :
\begin{equation}\label{J39}
\chi_{(\pm)}=\chi^{\;\prime}\left(\begin{array}{cccc}
\gamma_3{\bf 1}_N & 0 & 0 & 0 \\
0 & \gamma_3{\bf 1}_N & 0 & 0 \\
0 & 0 & \gamma_3{\bf 1}_N & 0 \\
0 & 0 & 0 & -\gamma_3{\bf 1}_N \end{array}\right)\;,\end{equation}
\item in case {\bf 2.b)} :
\begin{equation}\label{J40}
\chi_{(\pm)}=\chi^{\;\prime}\left(\begin{array}{cccc}
-\gamma_3{\bf 1}_N & 0 & 0 & 0 \\
0 & \gamma_3{\bf 1}_N & 0 & 0 \\
0 & 0 & \gamma_3{\bf 1}_N & 0 \\
0 & 0 & 0 & \gamma_3{\bf 1}_N \end{array}\right)\;,\end{equation}
\end{itemize}
\item[the real structure :] \({\bf J}=\left(\begin{array}{cc}
0 & {\bf J}_{(-)}\\ {\bf J}_{(+)} & 0 \end{array}\right)\), exchanges ${\cal H}_{(+)}$ and ${\cal H}_{(-)}$ and
\({\bf J}_{(\pm)}={\cal C}_1\otimes{\cal C}_2\chi_2\) yields
\begin{itemize}
\item in case {\bf 2.a)} :
\begin{equation}\label{J41}
{\bf J}_{{\bf a}(\pm)}=\chi^{\;\prime}\;\left(\begin{array}{cccc}
{\cal C}_{1}{\bf 1}_N & 0 & 0 & 0 \\
0 & 0 & {\cal C}_1{\bf 1}_N & 0 \\
0 & {\cal C}_1{\bf 1}_N & 0 & 0 \\
0 & 0 & 0 & -{\cal C}_1{\bf 1}_N \end{array}\right)\,{\cal K}\;,\end{equation}
\item in case {\bf 2.b)} :
\begin{equation}\label{J42}
{\bf J}_{{\bf b}(\pm)}=\chi^{\;\prime}\;\left(\begin{array}{cccc}
-{\cal C}_{1}{\bf 1}_N & 0 & 0 & 0 \\
0 & 0 & {\cal C}_1{\bf 1}_N & 0 \\
0 & {\cal C}_1{\bf 1}_N & 0 & 0 \\
0 & 0 & 0 & {\cal C}_1{\bf 1}_N \end{array}\right)\,{\cal K}\;,\end{equation}
\end{itemize}
\end{description}
\subsection{The "Real" Yang-Mills-Higgs action}\label{RealYMH}
The representations $\pi_{(\pm)}$ of (\ref{J35}) and $\pi^{o}_{(\pm)}$ of (\ref{J36}), with the Dirac operators of
(\ref{J37}) and (\ref{J38}), induce representations of $\Omega^\bullet({\cal A})$.\\
Let $F\in\Omega^{(1)}({\cal A})$, then, using the same techniques which led to (\ref{TRIP8}), we obtain in case {\bf 2.a)}:
\[
\pi_{(\pm)}(F)=
\left(\begin{array}{cccc}
-{\bf i}\;{\bf c}(\sigma^{(1)}_a){\bf 1}_{N} & 0 & 0 & 0\\
0 & -{\bf i}\;{\bf c}(\sigma^{(1)}_a){\bf 1}_{N} & 0 & \sigma^{(0)}_{ab}\;\gamma_3 A^+ \\
0 & 0 & -{\bf i}\;{\bf c}(\sigma^{(1)}_b){\bf 1}_{N}& 0 \\
0 & \sigma^{(0)}_{ba}\;\gamma_3 A & 0 & -{\bf i}\;{\bf c}(\sigma^{(1)}_b){\bf 1}_{N}
\end{array}\right)\;,\]
and in case {\bf 2.b)}:
\[
\pi_{(\pm)}(F)=
\left(\begin{array}{cccc}
-{\bf i}\;{\bf c}(\sigma^{(1)}_a){\bf 1}_{N} & 0 & \sigma^{(0)}_{ab}\;\gamma_3 A^+ & 0\\
0 & -{\bf i}\;{\bf c}(\sigma^{(1)}_a){\bf 1}_{N} & 0 & 0 \\
\sigma^{(0)}_{ba}\;\gamma_3 A & 0 & -{\bf i}\;{\bf c}(\sigma^{(1)}_b){\bf 1}_{N}& 0 \\
0 & 0 & 0 & -{\bf i}\;{\bf c}(\sigma^{(1)}_b){\bf 1}_{N}
\end{array}\right)\;,\]
where the differential forms $\sigma^{(k)}$'s are given in (\ref{one-form}).\\ Introducing the $2N\times 2N$ matrix
in case {\bf 2.a)} as \(M_{\bf (2.a)}=\left(\begin{array}{cc}0 & 0 \\ 0 & A\end{array}\right)\) and in case {\bf 2.b)} as \(M_{\bf 2.b)}=\left(\begin{array}{cc}A & 0 \\ 0 & 0\end{array}\right)\), we may also write:
\begin{equation}\label{J43}
\pi_{(\pm)}(F)=
\left(\begin{array}{cc}
-{\bf i}\;{\bf c}(\sigma^{(1)}_a){\bf 1}_{2N} &
\sigma^{(0)}_{ab}\;\gamma_3 M^+ \\
\sigma^{(0)}_{ba}\;\gamma_3 M &
-{\bf i}\;{\bf c}(\sigma^{(1)}_b){\bf 1}_{2N}
\end{array}\right)\;.\end{equation}
A universal two-form $G\in\Omega^{(2)}({\cal A})$ has representations $\pi_{(\pm)}(G)$ given by a similar expression as
in (\ref{TRIP9}). The unwanted differential ideal ${\cal J}$ is removed using the orthonality condition analogous to (\ref{Nojunk})
\begin{equation}\label{J44}
\rho_{aaa}^{(0)}-{1\over 2N}\rho_{aba}^{(0)}\;{\bf tr}\{M^+M\}=0\quad;\quad
\rho_{bbb}^{(0)}-{1\over 2N}\rho_{bab}^{(0)}\;{\bf tr}\{MM^+\}=0\;.\end{equation}
The representative of $G$ in $\Omega^{(2)}_D({\cal A})$, as in (\ref{TRIP15}), is given by :
\begin{equation}\label{J45}
\pi_{D(\pm)}(G)=\left(\begin{array}{cc}
\pi_{D(\pm)}(G)_{[aa]} & \pi_{D(\pm)}(G)_{[ab]}\\
\pi_{D(\pm)}(G)_{[ba]} & \pi_{D(\pm)}(G)_{[bb]}\end{array}\right)\;,\end{equation}
with
\begin{eqnarray*}
\pi_{D(\pm)}(G)_{[aa]}&=&-{\bf c}(\rho_{aaa}^{(2)})\otimes{\bf 1}_{2N}+ \rho_{aba}^{(0)}\otimes\Bigl[M^+M\Bigr]_{NT} \;,\\
\pi_{D(\pm)}(G)_{[ab]}&=&-{\bf i}\;{\bf c}(\rho_{ab}^{(1)})\gamma_{3(\pm)}\otimes M^+\;,\\
\pi_{D(\pm)}(G)_{[ba]}&=&-{\bf i}\;{\bf c}(\rho_{ba}^{(1)})\gamma_{3(\pm)}\otimes M\;,\\
\pi_{D(\pm)}(G)_{[bb]}&=&-{\bf c}(\rho_{bbb}^{(2)})\otimes{\bf 1}_{2N}+\rho_{bab}^{(0)}\otimes\Bigl[MM^+\Bigr]_{NT}\;,
\end{eqnarray*}
where the differential forms $\rho^{(k)}$ are given in (\ref{two-form}) and
\begin{eqnarray*}
\Bigl[M^+M\Bigr]_{NT}&=& M^+M-{1\over 2N}{\bf tr}\{M^+M\}\;,\\
\Bigl[M^+M\Bigr]_{NT}&=& MM^+-{1\over 2N}{\bf tr}\{MM^+\}\;.\end{eqnarray*}
The scalar product in $\Omega^{(2)}_D({\cal A})$ is the same for representatives in ${\cal H}_{(+)}$ or ${\cal H}_{(-)}$ and is given by a similar expression as (\ref{TRIP16}) :
\begin{eqnarray}\label{J46}
&&\langle \pi_{D(\pm)}(G);\pi_{D(\pm)}(G^\prime)\rangle_{2,D}=\nonumber\\
&&{1\over 2\pi}
\Biggl\{
2N\;\int_{S^2}\left({\rho^{(2)}_{aaa}}^*\wedge\star\rho^{\prime(2)}_{aaa}
+{\rho^{(2)}_{bbb}}^*\wedge\star\rho^{\prime(2)}_{bbb}\right)\nonumber\\
&&
+{\bf tr}\{A^+A\}\;
\int_{S^2}\left({\rho^{(1)}_{ab}}^*\wedge\star\rho^{\prime(1)}_{ab}
+
{\rho^{(1)}_{ba}}^*\wedge\star\rho^{\prime(1)}_{ba}\right)\nonumber\\
&&
+\left[{\bf tr}\{(A^+A)^2\}-{1\over 2N}({\bf tr}\{A^+A\})^2\right]\;
\int_{S^2}\left({\rho^{(0)}_{aba}}^*\wedge\star\rho^{\prime(0)}_{aba}
+
{\rho^{(0)}_{bab}}^*\wedge\star\rho^{\prime(0)}_{bab}\right)\Biggr\}\;.\nonumber\\
&&
\end{eqnarray}
From this expression of the scalar product, the Yang-Mills-Higgs action is essentially twice the action (\ref{YMH7}) obtained in section {\bf \ref{Yang-Mills-Higgs}} :
\begin{eqnarray}\label{J47}
&&{\bf S}_{YMH}(\nabla_D)={\lambda\over \pi}\biggl\{\nonumber\\
&& 2N\;\int_{S^2}\left({\bf tr}_{matrix}\Bigl\{(\!( F_a)\!)^+\wedge\star(\!( F_a)\!)\Bigr\}
+F_b^*\wedge\star F_b\right)\nonumber\\
&& +2\;{\bf tr}\{A^+A\}\;
\int_{S^2}\langle\nabla\eta_{ba}\vert\wedge\star\vert\nabla\eta_{ab}\rangle\nonumber\\
&& +\left[{\bf tr}\{(A^+A)^2\}-{1\over 2N}({\bf tr}\{A^+A\})^2\right]\;
\int_{S^2}\star\Bigl(2(\langle \eta_{ba}\vert \eta_{ab}\rangle-1)^2+1\Bigr)\biggr\}\;.\nonumber\\
&&\end{eqnarray}
\subsection{The "Real" covariant Dirac operator}
\label{RealMatter}
Matter\footnote{In this section we consider case {\bf 2.a)} only. At the end the final result
for the covariant Dirac operator in case {\bf 2.b)} will also be given} in this "real spectral
triple" approach is represented by states of the covariant Hilbert space
${\cal H}_{Cov}={\cal M}\otimes_{\cal A}{\cal H}\otimes_{\cal A}{\cal M}^*$. Fot the $S^0$-real spectral triple
${\cal H}={\cal H}_{(+)}\oplus{\cal H}_{(-)}$ so that also ${\cal H}_{Cov}$ splits in a sum of "particle" and
"antiparticle" Hilbert spaces ${\cal H}_{Cov}={\cal H}_{(+{\bf p})}\oplus{\cal H}_{(-{\bf p})}$ where each
${\cal H}_{(\pm{\bf p})}={\cal M}\otimes_{\cal A}{\cal H}_{(\pm)}\otimes_{\cal A}{\cal M}^*$ has typical elements
\(\Vert{\bf \Psi}_{(\pm{\bf p}}\rangle\!\rangle\). The projective module ${\cal M} ={\bf P} {\cal A}^2$ and its dual ${\cal M}^*$
were examined in section {\bf \ref{Module}} and in appendix {\bf \ref{ModCon}}. In bases $\{E_i\}$
and $\{E^j\}$ of the free modules ${\cal A}^2$ and ${{\cal A}^2}^*$, we may represent a state of
${\cal A}^2\otimes_{\cal A}{\cal H}_{(\pm)}\otimes_{\cal A}{{\cal A}^2}^*$ as
\(E_i\otimes_{\cal A}{\Psi_{(\pm)}}^i_{\;j}\otimes_{\cal A} E^j\). It is a state
\(\Vert{\bf \Psi}_{(\pm{\bf p}}\rangle\!\rangle\) of ${\cal H}_{(\pm{\bf p})}$ if ${\Psi_{(\pm)}}^i_{\;j}\in {\cal H}_{(\pm)}$
obeys
\begin{equation}\label{J48}
{\Psi_{(\pm)}}^i_{\;j}=\pi_{(\pm)}(P^i_k)\pi_{(\pm)}^o(P^\ell_j)\,{\Psi_{(\pm)}}^k_{\;\ell}\;.\end{equation}
Since $P^i_{\;k,a}(x)=\delta^i_k$ and $P^i_{\;k,b}(x)=\vert\nu(x)\rangle^i\langle\nu(x)\vert_k$
(cf. section {\bf \ref{Module}}) the vector ${\Psi_{(\pm)}}^i_{\;j}\in{\cal H}_{(\pm)}$ is represented
by the column vector
\begin{equation}\label{J49}
{\Psi_{(\pm)}}^i_{\;j}=\left(\begin{array}{c}
{(\!(\psi_{(\pm)aa})\!)}^i_{\;j}\\
{\vert\psi_{(\pm)ab}\rangle}^i\langle\nu\vert_j\\ \vert\nu\rangle^i{\langle\psi_{(\pm)ba}\vert}_j\\
\vert\nu\rangle^i\;{\psi_{(\pm)bb}}\;\langle\nu\vert_j\end{array}\right)\;,\end{equation}
where $(\!(\psi_{(\pm)aa})\!)$ is a quadruplet and
$\psi_{(\pm)bb}=\langle\nu\vert(\!(\psi_{(\pm)bb})\!)\vert\nu\rangle$ a singlet of Pensov spinor
fields of spin weight $\pm s$, while
$\vert\psi_{(\pm)ab}\rangle=(\!(\psi_{(\pm)ab})\!)\vert\nu\rangle$, respectively
$\langle\psi_{(\pm)ba}\vert=\langle\nu\vert(\!(\psi_{(\pm)ba})\!)$, are doublets of Pensov
spinors of weight $(\pm s)-n/2$, respectively $(\pm s)+n/2$.\\ The covariant real structure
${\bf J}_{Cov}$, as defined in appendix {\bf \ref{RealCov}}, acts on $\Vert{\bf \Psi}_{Cov}\rangle\!\rangle$ as
\begin{equation}\label{J50}
{\left({\bf J}_{Cov}\Vert{\bf \Psi}_{Cov}\rangle\!\rangle\right)_{(\pm)}}^i_{\;j}=\chi^{\;\prime}
\left(\begin{array}{c}
\delta^{i\bar\ell}\left({\cal C}_1{\cal K}(\!(\psi_{(\mp)aa})\!)^k_{\;\ell}\right)\delta_{\bar kj}\\
\delta^{i\bar\ell}\left({\cal C}_1{\cal K}\langle\psi_{(\mp)ba}\vert_\ell\right)\langle\nu\vert_j\\
\vert\nu\rangle^i\left({\cal C}_1{\cal K}\vert\psi_{(\mp)ab}\rangle^k\right)\delta_{\bar kj}\\
-\vert\nu\rangle^i\left({\cal C}_1{\cal K}\psi_{(\mp)bb}\right)\langle\nu\vert_j\end{array}\right)\;.\end{equation}
\newpage
\noindent
The covariant Dirac operator, defined in (\ref{Cov10}), is also block diagonal : ${\cal D}_\nabla={\cal D}_{(+{\bf p})}\oplus{\cal D}_{(-{\bf p})}$ and is given by :
\begin{eqnarray}\label{J51}
&&{\cal D}_{(\pm{\bf p})}\biggl\{(E_i P^i_{\;k})\otimes_{\cal A}{\psi_{(\pm)}}^k_{\;\ell}\otimes_{\cal A} (P^\ell_{\;j}E^j)\biggr\}
=\nonumber\\
&&
E_i\otimes_{\cal A}\pi_{(\pm)}\left((\!( A)\!)^i_{\;k}\right){\psi_{(\pm)}}^k_{\;j}
\otimes_{\cal A} E^j\nonumber\\
&&+E_i\otimes_{\cal A}
\pi_{(\pm)}(P^i_{\;k})\pi_{(\pm)}^o(P^\ell_j){\cal D}_{(\pm)}{\psi_{(\pm)}}^k_{\;\ell}
\otimes_{\cal A} E^j\nonumber\\
&&+\epsilon^{\;\prime}E_i\otimes_{\cal A}
\pi_{(\pm)}^o\left((\!( A)\!)^\ell_j\right){\psi_{(\pm)}}^i_{\;\ell}
\otimes_{\cal A} E^j\;,\end{eqnarray}
where $(\!( A)\!)$ is the $2\times 2$ matrix of universal one-forms given in (\ref{PMod6}).\\
It is represented by the matrix valued differential one- and zero-forms given in terms of \((\!(\alpha_a)\!)\), \(\alpha_b\), \(\vert\Phi_{ab}\rangle\) and \(\langle\Phi_{ba}\vert\), defined in
(\ref{YMH1}) by :
\begin{eqnarray}\label{J52}
(\!( A_{a})\!)&=&-{\bf i}\gamma^r(\!(\alpha_{a,r})\!)\;,\nonumber\\
(\!( A_{b})\!)&=&-{\bf i}\gamma^r\alpha_{b,r}\vert\nu\rangle\langle\nu\vert\;,\nonumber\\
(\!( A_{ab})\!)&=&\vert\Phi_{ab}\rangle\langle\nu\vert\;\gamma_{3}\;,\nonumber\\
(\!( A_{ba})\!)&=&\vert\nu\rangle\langle\Phi_{ba}\vert\;\gamma_{3}\;.
\end{eqnarray}
We obtain :
\begin{equation}\label{J53}
\pi_{(\pm)}((\!( A)\!))=
\left(\begin{array}{cccc}
(\!( A_{a})\!){\bf 1}_N & 0 & 0 & 0 \\
0 & (\!( A_{a})\!){\bf 1}_N & 0 & (\!( A_{ab})\!) {\bf A}^+ \\
0 & 0 & (\!( A_{b})\!){\bf 1}_N & 0 \\
0 & (\!( A_{ba})\!) {\bf A} & 0 & (\!( A_{b})\!){\bf 1}_N
\end{array}\right)\,.\end{equation}
The $\pi_{(\pm)}^o$ representative of $(\!( A)\!)$ is computed as :
\begin{equation}\label{J54}
\epsilon^{\;\prime}\pi_{(\pm)}^o((\!( A)\!))=
\left(\begin{array}{cccc}
(\!( A_{a})\!){\bf 1}_N & 0 & 0 & 0 \\
0 & (\!( A_{b})\!){\bf 1}_N & 0 & 0 \\
0 & 0 & (\!( A_{a})\!){\bf 1}_N & -(\!( A_{ba})\!) {\bf B}^+ \\
0 & 0 & -(\!( A_{ab})\!) {\bf B} & (\!( A_{b})\!){\bf 1}_N
\end{array}\right)\;.\end{equation}
\newpage
\noindent
Substituting (\ref{J53}) and (\ref{J54}) in (\ref{J51}), we obtain :
\begin{eqnarray}
(\!({\cal D}_\nabla\Vert{\bf \Psi}_{Cov}\rangle\!\rangle_{(\pm)aa})\!)&=&
{\cal D}_{1(\pm)}(\!(\psi_{(\pm)aa})\!) -{\bf i}\;\gamma^r\left[(\!(\alpha_{a,r})\!),(\!(\psi_{(\pm)aa})\!)\right]\;,\nonumber\\
\vert{\cal D}_\nabla\Vert{\bf \Psi}_{Cov}\rangle\!\rangle_{(\pm)ab}\rangle&=&
{\cal D}^{(-n/2)}_{1(\pm)} \vert\psi_{(\pm)ab}\rangle+\vert\eta_{ab}\rangle\gamma_{3}{\bf A}^+\psi_{(\pm)bb}\nonumber\\
&&-{\bf i}\;\gamma^r\left((\!(\alpha_{a,r})\!)\vert\psi_{(\pm)ab}\rangle
-\vert\psi_{(\pm)ab}\rangle\alpha_{b,r}\right)\;;\nonumber\\
\langle{\cal D}_\nabla\Vert{\bf \Psi}_{Cov}\rangle\!\rangle_{(\pm)ba}\vert&=&
{\cal D}^{(+n/2)}_{1(\pm)}\langle\psi_{(\pm)ba}\vert+\langle\eta_{ba}\vert\gamma_{3}{\bf B}^+\psi_{(\pm)bb}\nonumber\\
&&-{\bf i}\;\gamma^r\left(\alpha_{b,r}\langle\psi_{(\pm)ba}\vert
-\langle\psi_{(\pm)ba}\vert(\!(\alpha_{a,r})\!)\right)\;,\nonumber\\
{\cal D}_\nabla\Vert{\bf \Psi}_{Cov}\rangle\!\rangle_{(\pm)bb}&=&
{\cal D}_{1(\pm)}\psi_{(\pm)bb}\nonumber\\
&&+\gamma_{3}{\bf A}\langle\eta_{ba}\vert\psi_{(\pm)ab}\rangle +
\gamma_{3}{\bf B}\langle\psi_{(\pm)ba}\vert\eta_{ab}\rangle\;.\nonumber\\
&&\label{J55}
\end{eqnarray}
In case {\bf 2.b)} a similar result is obtained :
\begin{eqnarray}
(\!({\cal D}_\nabla\Vert{\bf \Psi}_{Cov}\rangle\!\rangle_{(\pm)aa})\!)&=&
{\cal D}_{1(\pm)}(\!(\psi_{(\pm)aa})\!) -{\bf i}\;\gamma^r\left[(\!(\alpha_{a,r})\!),(\!(\psi_{(\pm)aa})\!)\right]\nonumber\\
&&+\gamma_{3}{\bf A}^+\;\vert\eta_{ab}\rangle\langle\psi_{(\pm)ba}\vert
+\gamma_{3}{\bf B}^+ \;\vert\psi_{(\pm)ab}\rangle\langle\eta_{ba}\vert\;, \nonumber\\
\vert{\cal D}_\nabla\Vert{\bf \Psi}_{Cov}\rangle\!\rangle_{(\pm)ab}\rangle&=&
{\cal D}^{(-n/2)}_{1(\pm)}\vert\psi_{(\pm)ab}\rangle
+\gamma_{3}{\bf B}\; (\!(\psi_{(\pm)aa})\!)\vert\eta_{ab}\rangle\;,\nonumber\\
&&-{\bf i}\;\gamma^r\left((\!(\alpha_{a,r})\!)\vert\psi_{(\pm)ab}\rangle
-\vert\psi_{(\pm)ab}\rangle\alpha_{b,r}\right)\;,\nonumber\\
\langle{\cal D}_\nabla\Vert{\bf \Psi}_{Cov}\rangle\!\rangle_{(\pm)ba}\vert&=&
{\cal D}^{(+n/2)}_{1(\pm)}\langle\psi_{(\pm)ba}\vert
+\gamma_{3}{\bf A} \;\langle\eta_{ba}\vert(\!(\psi_{(\pm)aa})\!)\;,\nonumber\\
&&-{\bf i}\;\gamma^r\left(\alpha_{b}\langle\psi_{(\pm)ba}\vert-\langle\psi_{(\pm)ba}\vert(\!(\alpha_{a,r})\!)\right)\nonumber\\
\left({\cal D}_\nabla\Vert{\bf \Psi}_{Cov}\rangle\!\rangle\right)_{(\pm)bb}&=&
{\cal D}_{1(\pm)}\psi_{(\pm)bb}\;.\nonumber\\
&&\label{J56}
\end{eqnarray}
The difference is that here the Higgs field interact with the quadruplet $(\!(\psi_{(\pm)aa})\!)$, while in case {\bf 2.a)} it interacts with the singlet $\psi_{(\pm)bb}$ and it is this interaction that gives masses to the particles.\\ The Dirac operators,
${\cal D}^{(-n/2)}_{1(\pm)}={\cal D}_{(\pm s-n/2)}$ and ${\cal D}^{(+n/2)}_{1(\pm)}={\cal D}_{(\pm s+n/2)}$,
acting on Pensov spinor fields of spin weight $\pm s+n/2$ or $\pm s-n/2$, arise from
\[\Bigl({\cal D}_{1(\pm)}(\vert\psi_{(\pm)ab}\rangle\langle\nu\vert)\Bigr)\vert\nu\rangle=
{\cal D}^{(-n/2)}_{1(\pm)}\vert\psi_{(\pm)ab}\rangle\;,\]
\[\langle\nu\vert\Bigl({\cal D}_{1(\pm)}(\vert\nu\rangle\langle\psi_{(\pm)ba}\vert)\Bigr)=
{\cal D}^{(+n/2)}_{1(\pm)}\langle\psi_{(\pm)ba}\vert\;,\]
where $(\!( P_b)\!)=\vert\nu\rangle\langle\nu\vert$ is the representative, chosen in (\ref{PMod3}), of the homotopy class $[n]\in\pi_2(S^2)$.
The induced contribution of the "magnetic monopole" is hidden in this modification of the Dirac operator.\\ The Higgs doublets of Pensov fields of weight $\mp \;n/2$, $\vert\eta_{ab}\rangle$ and $\langle\eta_{ba}\vert$ were defined in (\ref{YMH3}).\\
A suitable action of the matter field would be
\begin{equation}\label{J59}
{\bf S}_{\bf Mat}(\Vert{\bf \Psi}_{Cov}\rangle\!\rangle,\nabla_D)=
\Bigl(\Vert{\bf \Psi}_{Cov}\rangle\!\rangle\;;\;{\bf D}_\nabla\Vert{\bf \Psi}_{Cov}\rangle\!\rangle\Bigr)\;.\end{equation}
But, if we aim for a theory admitting only chiral matter, i.e. if we restrict the Hilbert space to those vectors obeying say
\begin{equation}\label{J60}
{\bf \Gamma}\Vert{\bf \Psi}_{Cov}\rangle\!\rangle=+\Vert{\bf \Psi}_{Cov}\rangle\!\rangle\;,\end{equation}
then the above action vanishes identically. Another choice for the action would be
\begin{equation}\label{J61}
{\bf S}_{\bf Chiral}(\Vert{\bf \Psi}_{Cov}\rangle\!\rangle,\nabla_D)=
\Bigl({\bf J}_{Cov}\Vert{\bf \Psi}_{Cov}\rangle\!\rangle\;;\;{\bf D}_\nabla\Vert{\bf \Psi}_{Cov}\rangle\!\rangle\Bigr)\;.\end{equation}
It is easy to show that this action does not vanish identically if
\begin{equation}\label{J62}
\epsilon^{\;\prime\prime}=-1\;,\end{equation}
\begin{equation}\label{J63}
\epsilon\;\epsilon^{\;\prime}=+1\;.\end{equation}
In two dimensions Connes' sign table obeys the first but not the second
condition\footnote{In four dimensions it is the first condition that is not satisfied.}. It should
however be stressed that Connes' sign table, with its modulo eight periodicity, comes from
representation theory of the real Clifford algebras and if we restrict our (generalized) spinors to
Weyl spinors, we loose the Clifford {\it algebra} representation and the sign tables ceases to be
mandatory. We could then go back to {\bf \ref{Prodtriple}} and
\begin{itemize}
\item with ${\bf J}={\bf J}_1\otimes{\bf J}_2$ require (\ref{J32}) to hold with
$\epsilon_1^{\;\prime}=-1\epsilon_1$ or,
\item with ${\bf J}={\bf J}_1\otimes({\bf J}_2\chi_2)$ require (\ref{J34}) to hold with
$\epsilon_1^{\;\prime}=+1=\epsilon_1$.
\end{itemize}
\newpage
\section{Conclusions and Outlook}
\label{Out}
The Connes-Lott model over the two-sphere, with ${\bf C}\oplus{\bf C}$ as discrete algebra, has been
generalized such as to allow for a nontrivial topological structure. The basic Hilbert space
${\cal H}_{(s)}$ was made of Pensov spinors which can be interpreted as usual spinors inteacting with a
Dirac monopole "inside" the sphere. Covariantisation of the Hilbert space
${\cal H}_{(s)}\otimes{\cal H}_{dis}$ with a nontrivial projective module ${\cal M}$ induced a "spin" change in
certain matter fields so that we obtained singlets and doublets of different spin content. The Higgs
fields also acquired a nontrivial topology since they are no longer ordinary functions on
the sphere, but rather Pensov scalars i.e. sections of nontrivial line bundles over the sphere.\\
A real spectral triple has also been constructed essentially through the doubling of the Pensov
spinors so that the Hilbert space of the continuum spectral triple
became ${\cal H}_1={\cal H}_{(s)}\oplus{\cal H}_{(-s)}$. The discrete spectral triple had also to be extended in
order that the first order condition could be met.
In contrast with the standard noncommutative geometry model of the standard model, in our model the continuum spectral triple has an $S^0$-real structure while the discrete spectral triple has not. Some physical plausability arguments for this were given in section {\bf \ref{RPDtriple}}. It was also shown that the covariantisation of the real spectral triple with the nontrivial ${\cal M}$ allows the abelian gauge fields to survive, while they are slain if covaraiantisation is done with a trivial module. Finally a possibility of solution to the the problem of a non vanishing action of chiral matter has been indicated, paying the price of using a complex action.\\
If we address the quantisation problem in a path integral formalism, let us first recall that we have Higgs fields which are Pensov scalars of a ${\cal P}^{(s_1)}$ and matter fields Pensov spinors of type ${\cal H}_{(s_2)}$. It is thus tempting to assume that the Higgs fields should be even Grassmann variables is $s_1$ is integer valued and odd Grassmann variables if it is half-integer. In the same vein the matter fields should be odd or even Grassmann variables if $s_2$ is integer or half integer valued. A thorough examination of this issue is however beyond the scope of this work.
\newpage
|
\section{Introduction}
While the study of localization of classical and quantum waves in
random disordered media has been well understood
\cite{Anderson,John,EcoSou,Sou}
, recently, the wave
propagation in amplifying random media has been pursued intensively
\cite{experi,Letokhov,PR,Zhang,dual,Zyuzin1,Zyuzin2,tightbind,India}.
Some interesting results have been predicted, such
as, the localization length of a random medium with gain \cite{Zhang},
the sharpness of back scattering
coherent peak \cite{experi,Zyuzin1}, the dual symmetry of
absorption and amplification
\cite{dual}, the critical size of the system \cite{Letokhov,Zhang},
and the probability distribution of
reflection \cite{PR}. Numerically, two kinds of models are studied, one is
the electronic tight binding model \cite{tightbind,India}, the
other is the many-layered model of classical waves\cite{Zhang}
. Theoretically, a lot of methods are used to get these results, such as
the diffusion theory \cite{Letokhov,Zyuzin2} and the transmission
matrix method \cite{PR}.
Most of these studies are for homogeneously random systems which are
generated by introducing the disorder into the continuous system, and the
medium parameter, such as the dielectric constant, is assumed to
vary in a continuous way \cite{experi,Zhang}.
But the periodically correlated random systems which are generated by
introducing the disorder into a periodic system, such as a
photonic-band-structure , have not been studied adequately.
\par
With gain, will such random systems with periodic background
behave similar as the homogeneously random system? Both experimentally and
theoretically, the study of such system is very important in understanding
the propagation of light in random media. These type of
photonic-band-structure
systems are widely used in experiments \cite{EcoSou,Yariv}.
Theoretically, just as S.John \cite{EcoSou}
argued, the localization of a photon is from a subtle interplay
between order and disorder. For the periodically correlated
random systems with gain, the periodic background plays the order role,
and now its interplay with not only disorder but also with gain should
be a very interesting new topic.
\par
In this paper we address both the electronic tight binding model and
the many-layered model of classical waves. We first compare the numerical
results of periodic amplifying system with what
we can predict theoretically by the transfer matrix method. It's
surprising to get most of the universal
properties, such as critical length and exponential decay of
transmission, of
homogeneously random system \cite{Zhang} from a
periodic system too.
With the help of some theoretical arguments and numerical results,
we suggest that the length ${\xi_1}=|1/Im(K)|$, where $K$ is the Bloch
vector in a periodic system with gain, to replace the gain length
${l_g}=|1/Im(k)|$ introduced in \cite{Zhang}. This is more
reasonable
since the correlated scatterers in a periodic system can make
the paths of wave propagation much longer in the system. We also
think that it is actually the Bloch wave instead of the plane wave which
propagates in the system.
Then we introduce disorder into these periodic systems and calculate
their properties. Our numerical simulations for
both
models show that periodically correlated random systems give similar
behaviors
as that of the homogeneously random systems studied previously.
But in some cases, we get
new
interesting results for the localization length $\xi$, the critical
length $L_c$ and the probability distribution $P(R)$ of reflection. All
these new results are related to the periodic background of such systems.
We also examine the results of the transmission coefficient $T$ for $short$
$(L<{L_c})$ systems. Our
numerical results show that the formula of the transmission coefficient of
media with
$absorption$ can be generalized to the transmission coefficient of short
systems
with $gain$, if we replace the gain length $l_g$ (or $\xi_1$) with
the negative of the absorption length $-l_a$ in the formula.
To explain our new results of the critical length $L_c$, we compare the two
basic theories for obtaining the critical length, the Letokhov theory
\cite{Letokhov}
and the Lamb theory \cite{LaserPhys}, and we get some new
theoretical results of critical length which are in good agreement with our
numerical results.
The behavior of the distribution of the probability of reflection $P(R)$ is
much
more complex than the theoretical prediction of homogeneously random
system \cite{PR}. We find that the periodic background influences strongly
the general behavior of $P(R)$.
\par
The paper is organized as follows. In Sec.II we introduce the two
theoretical models we are studying. The results for the periodic systems
with gain are presented in Sec.III, while in Sec.IV the results for the
random systems with gain are given. Also in Sec.IV we present our new
theoretical and numerical results for the critical length $L_c$. In Sec.V
the results for the probability distribution of reflection coefficient
$R$ for both models are presented. Finally, Sec.VI is devoted to a
discussion of our results and give some conclusions.
\section{Theoretical Models}
\subsection{ Many-layered model of classical wave}
Our periodic many-layered model of classical wave consists of two types of
layers with
dielectric constant ${\varepsilon_1}={\varepsilon_0}-i\varepsilon''$ and
${\varepsilon_2}=\chi{\varepsilon_0}-i\varepsilon''$ and thicknesses
$a=95nm$ and ${b_0}=120nm$ respectively, where the negative part of
dielectric constant, i.e. $\varepsilon''>0$, denotes the homogeneous
amplification of the field. We have tried a lot values for $\chi$, such
as 1.5, 2, 3, 5, 6, and
get no essential difference in our results for different values. In this
paper we choose
$\chi=2$, i.e. $Re(\varepsilon_1)=\varepsilon_0=1$ and
$Re(\varepsilon_2)=2\varepsilon_0=2$. The system has L cells. Each cell is
composed by two layers with dielectric constant ${\varepsilon_1}$ and
${\varepsilon_2}$ respectively.
Without gain, we obtain that the wavelength range of the second band of
this
periodic system is from 247 nm to 482.6 nm ( the first band has a range from
592 nm to infinite). So we
choose the wavelength 360 nm to represent band center,
the wavelength 420nm as a general case, and the wavelength 470 nm to
represent the band edge.
To introduce disorder, we
choose the width of second layer of the $n$th cell to be random variable
${b}_n={b}_0(1+W\gamma)$, where $W$ describes the
strength of randomness and $\gamma$ is a random number between $(-0.5,
0.5)$.
The whole system is embedded in a homogeneous infinite material with
dielectric constant equal to ${\varepsilon_0}$.
For the 1D case, the time-independent Maxwell equation can be written as:
\begin{eqnarray}
\frac{{\partial^2} E(z)}{\partial {z^2}} +
{\omega^2\over{c^2}}\varepsilon(z)E(z)=0 \label{Maxw}
\end{eqnarray}
Suppose that in the medium with dielectric constant $\varepsilon_1$ and the
medium
with dielectric constant $\varepsilon_2$, the electric field \cite{Zhang} is
given by the following expressions:
\begin{eqnarray}
E_{1n}(z) &=& {A_n}{e^{ik(z-{z_n})}}+{B_n}{e^{-ik(z-{z_n})}} \nonumber \\
E_{2n}(z) &=& {C_n}{e^{ik(z-{z_n})}}+{D_n}{e^{-ik(z-{z_n})}} \label{feild}
\end{eqnarray}
Using the appropriate boundary condition (continuity of the electric
field $E$ and of the derivative of $E$ at the interface), we obtain that:
\begin{eqnarray}
\left( \begin{array}{c} {A_{n-1}} \\ {B_{n-1}} \end{array} \right) =
\left( {M_n} \right)
\left( \begin{array}{c} {A_{n}} \\ {B_{n}} \end{array} \right)
\end{eqnarray}
where
\begin{eqnarray}
\left( {M_n} \right) =
\left( \begin{array}{clcr}
{e^{-ika}}(cos(q{b_n})-{i\over2}
({\frac{k}{q}}+{\frac{q}{k}})sin(q{b_n}) &
-{i\over2}{e^{-ika}}({\frac{k}{q}}-{\frac{q}{k}})sin(q{b_n}) \\
{i\over2}{e^{ika}}({\frac{k}{q}}-{\frac{q}{k}})sin(q{b_n}) &
{e^{ika}}(cos(q{b_n})+{i\over2}
({\frac{k}{q}}+{\frac{q}{k}})sin(q{b_n})
\end{array} \right) \label{tranM}
\end{eqnarray}
where $ k={\frac{\omega}{c}}\sqrt{\varepsilon_1} $ and
$q={\frac{\omega}{c}} \sqrt{\varepsilon_2} $.
From the product of these matrices, $M(L)={\prod_{1}^L}{M_n} $, we can
obtain the transmission and reflection amplitudes of the sample,
$t(L)=\frac{1}{M_{11}}$ and $r(L)=\frac{M_{21}}{M_{11}}$.
For each set of parameters $(L, W, {\varepsilon''})$, the reflection
coefficient $R=|r|^2$ and the transmission coefficient $T=|t|^2$
are obtained from a large number of random configurations. We have used
10,000 configurations to calculate the different average values of
$R$ and $T$, and 1,000,000 configurations to obtain
$P(R)$.
Our numerical results
show that the localization length for a system without gain behaves
$\xi_0 \propto
1/{W^2}$ for this model, and are in agreement with previous workers.
\subsection{ Electronic tight-binding model}
\par
For the electronic tight-binding model, the wave equation can be written as:
\begin{eqnarray}
\left( \begin{array}{c} {\phi_{n+1}} \\ {\phi_n} \end{array} \right)
=\left( {M_n} \right) \left( \begin{array}{c} {\phi_n} \\ {\phi_{n-1}}
\end{array} \right)
\end{eqnarray}
where
\begin{eqnarray}\label{tim}
\left( {M_n} \right)=
\left( \begin{array}{clcr} E-{\epsilon_n} & -1 \\
1 & 0 \end{array} \right)
\end{eqnarray}
${\epsilon_n}=W{\gamma}-i\eta$, where W describes the strength of
randomness, $\gamma$ is a random number between $(-0.5, 0.5)$ ,
$\eta>0$ corresponds to amplification and $\phi_n$ is
the wave function at site $n$. The length L of the system is
the total lattice number of the system. The system is embedded
in two identical semi-infinite perfect leads on either side. For the left
and the right sides, we have ${\phi_0}=1+r(L)$ and
${\phi_{L+1}}=t(L){e^{ik(L+1)}}$. We can obtain reflection amplitude $r(L)$
and transmission amplitude $t(L)$ by the products of matrices,
$M(L)={\prod_{1}^L}{M_n} $. \begin{eqnarray}
t(L)=\frac{-2isin(k)}{{M_{11}}{e^{-ik}}+{M_{12}}-{M_{22}}{e^{ik}}-{M_{21}}}
{e^{-ik(L+1)}} \nonumber \\
r(L)=\frac{{M_{21}}{e^{ik}}+{M_{22}}-{M_{11}}-{M_{12}}{e^{-ik}}}
{{M_{11}}{e^{-ik}}+{M_{12}}-{M_{21}}-{M_{22}}{e^{ik}}}
{e^{-ik}} \label{tTR}
\end{eqnarray}
where $ k=arccos(\frac{E}{2})$.
When $W=0$ and without gain, the model is a periodic one with only one
band spanning in energy between -2 and 2. Notice that the hopping matrix
elements in Eq.(\ref{tim}) are equal to one, which is our unit of energy.
So we choose
$E=0$ to represent band center, $E=1$ as a general case, $E=1.8$ to
represent band edge.
Similar as the many-layered
model, for each set of parameters $(L, W, \eta)$, 10,000
random configurations were used to obtain a
average value of $R$ and $T$, and one million random configurations for
$P(R)$.
Theoretical and numerical results give that the localization
length for a system without gain behaves $\xi_0 \propto 1/{W^2}$, in
agreement with previous workers.
\section{Periodic Systems}
Almost all the properties of the periodic systems of both the many-layered
model and the tight-binding model can be predicted theoretically.
\subsection{Classical many-layered model}
\par
For long systems ($L\gg {L_c^0}$) of the many-layered model we have that:
\begin{eqnarray}
\lim_{L \rightarrow \infty} \frac{\partial \ln{T}}{\partial
L}&=&
-{{\xi_1}^{-1}} \nonumber \\
&=&2Im(K) \propto \varepsilon''
\end{eqnarray}
where $K$ is the Bloch vector which is a complex number now, and satisfies
$cosK=cos(ka)cos(q{b_0})
-{\frac{1}{2}}(\frac{k}{q}+\frac{q}{k})sin(ka)sin(q{b_0})$.
Because $ Im(K)<0$, the transmission coefficient $T$ is decaying
exponentially for a long system.
For a short system, $ L < {L_c^0} $, we have:
\begin{eqnarray}\label{pct}
\frac{\partial \ln{T}}{\partial L} =
1/{\xi_1'} & \simeq & 2|C||Im(K)| \nonumber \\
& \simeq & 2|Im(K)|
\end{eqnarray}
Where $C=-[sin(ka)cos(q{b_0})+cos(ka)sin(q{b_0})]/[2sin(K)] $, and $|C|$ is
larger than but very close to 1 when wavelength is at the band center,
and become bigger when wavelength approached the band edge.
So the slope of $\ln{T}$ vs $L$ for a short periodic system is almost same
as the negative value of the slope for the long system. The slopes of
$\ln{T}$ at both sides of the maximum
are approximately symmetric.
In Fig.~1a, we can see that, when $L < L_c^0$, $T$ increases vs L with
the slope $1/\xi_1'$, and get to a maximum at $ L_c^0$ and decays
exponentially when $L >L_c^0$ with the slope $1/\xi_1$.
From the behavior of the theoretical expressions of $T$ or $R$ , when
$T$ or $R$ goes to infinite, we can obtain analytically that $L_c^0$ is
given as:
\begin{eqnarray}\label{LC0}
{L_c^0} \simeq {\xi_1}\ln {\left(\frac{|C|+1}{|C|-1}\right)}
\end{eqnarray}
where $C$ is same as defined above in Eq.(\ref{pct}), and $|C|$ is
close to one. From the property of $|C|$ discussed above, we can see
that
$ {L_c^0} >> {\xi_1}$ at the band center , and becomes smaller when
wavelength approaches
the band edge. The value of $|C|$ is almost independent of gain, so
${L_c^0}$
is parallel to $1/\varepsilon''$ or $\xi_1$. We have shown that
$L_c^0\varepsilon''$ is almost a constant for a given wavelength, and
our numerical results agree very well with the theoretical prediction.
The reflection coefficient gets to a maximum value at $L_c^0$ too, and
fluctuates a
lot with the size $L$ of the system. When $L$ approaches infinity, $R$
reaches a saturated value. The saturated value of $R$ is given by:
\begin{eqnarray}
\lim_{L \rightarrow \infty} R =R_0 \simeq
\frac{{|(\frac{k}{q}-\frac{q}{k})sin(qb)|}^2}{{|sin(K)(|C|-1)|}^2}
\end{eqnarray}
So $R_0$ is almost independent of gain or $\xi_1$. Fig.~2 shows that
indeed $R$
increase when $L<L_c^0$, gets to its maximum value and fluctuate
violently at
$L_c^0$, then approaches a saturated value which is almost independent of gain.
\subsection{ Electronic tight-binding model}
\par
For the electronic tight-binding model, when $E=0$, the $\ln{T}$ for the long
system can be obtained by the use of Eq.(7):
\begin{eqnarray}
\lim_{L \rightarrow \infty} \frac{\partial \ln{T}}{\partial L}
=-{1/\xi_1} \simeq -\eta
\end{eqnarray}
\par
Similarly as the classical many-layered model, for short system of
tight-binding model we have:
\begin{eqnarray}
\frac{\partial \ln{T}}{\partial L}=1/{\xi_1'} \simeq \eta
\end{eqnarray}
So the slope symmetry of $\ln{T}$ at both sides of $L_c^0$ still exists.
In Fig.~1b, we can see a similar behavior as in Fig.~1a, when $L<L_c^0$,
$\ln{T}$ change vs $L$ with the slope of $1/\xi_1'$ and gets maximum at
$L_c^0$,
then it begin to decay exponentially with a slope of $1/\xi_1$.
Assuming that the theoretical expression of $T$, given by Eq.(7), is
infinite, we can obtain that $L_c^0$ is given by:
\begin{eqnarray}
{L_c^0} \simeq {\frac{2}{\eta}}(\ln{4}-\ln{\eta}) \simeq 2{\xi_1}\ln{(4{\xi_1})}
\end{eqnarray}
We also shown that ${L_c^0}\eta+2\ln{\eta}$ vs $\eta$, for $E=0$, is a
constant for different gain and indeed
find out that the theoretical prediction given by Eq.(14) agree very well
with the numerical results.
The reflection coefficient $R$ approaches a saturated value as L goes
to infinite, but the
saturated value of $R_0$ is not a constant independent of the gain as in the
case of
classical many-layered model. This is clearly seen in Fig.~3 where we
plot $\ln{R}$ vs L. Notice that the
$\ln{R}$ curves increase vs $L$ when $L<L_c^0$, get to a maximum at
$L_c^0$, and then approach a saturated value when L goes to infinity.
Similar results were obtained for $E \neq 0$.
\section{Random Systems}
In Fig.~4, we give the general behavior of
average value $<\ln{T}>$ vs $L$ for both models.
We can see the different behaviors for $L<L_c$ and $L>L_c$. When
$1/\xi_1> 1/\xi_0$ and $L<L_c$,
$<\ln{T}>$ increase vs L from origin with a slope which is defined as
$1/\xi'$ and when
$L>L_c$, $<\ln{T}>$ decrease vs L with a slope $-1/\xi$. But when $1/\xi_1<
1/\xi_0$, $<\ln{T}>$ will decrease monotonically, at first with the slope
$1/\xi_1=-1/|\xi_1|$, at $L_c$, there are a turning point and slope
changes to $-1/\xi$. We
will study the values of $\xi'$ , $\xi$ and $L_c$ in this section.
It was first suggested by Zhang \cite{Zhang} that the localization length
$\xi$ of a long random system with gain will become smaller than the
localization length $\xi_0$ of the random system without gain. In
particular he suggested that:
\begin{eqnarray}
\frac{1}{\xi}=\lim_{L \rightarrow \infty}
\frac{\partial \ln{T}}{\partial L}=\frac{1}{\xi_0}+\frac{1}{\xi_1}
\label{locleng}
\end{eqnarray}
where ${\xi_0}$ is the localization length of the system without gain,
${l_g}$ is replaced by ${\xi_1}$ in the original formula of Zhang
because of the periodic background of our systems.
We have numerically calculated $1/\xi$ for different cases of
disorder, gain and frequency(energy), and compare it with
$1/{\xi_0}+1/{\xi_1}$, as shown in Fig.~5a and 5b for
the many-layers model and the tight-binding model respectively. For most
of the cases, Eq.(\ref{locleng}) is a very good formula. Only when
wavelength is on band edge $and$ when both gain and randomness are very
strong, we can see that the numerical results deviate
from the theoretical prediction, which is the solid line in both Fig.~5a
and~5b.
\par
For short system($L<L_c$), the behavior of $<\ln{T}>$ vs $L$ is quite
different from that of long system as shown in Fig.~4. Freilikher
et.al. and Rammal and Doucot \cite{short}
obtained that the transmission coefficient of a random system with
$absorption$ is given by:
\begin{eqnarray}
<\ln{T}>=(-\frac{1}{l_a}-\frac{1}{\xi_0})L \label{absorb0}
\end{eqnarray}
where ${l_a}$ is the absorption length and ${\xi_0}$ is the localization
length. For a medium
with $gain$, can we just substitute the $-{l_a}^{-1}$ with ${l_g}^{-1}$
in the Eq.(\ref{absorb0}) to get
following equation?
\begin{eqnarray}
\frac{<\ln{T}>}{L}=\frac{1}{\xi'}=(\frac{1}{l_g}-\frac{1}{\xi_0}) \label{ab}
\end{eqnarray}
Because of the periodic background of our models, we use $\xi_1$ to replace
$l_g$ in our calculations.
So far there is no independent verification for this conclusion. After
substituting $\xi_1$ for $l_g$, our numerical results show
that Eq.(\ref{ab}) is correct for $short$ systems with $gain$ for both
models.
When the strength of the disorder is a constant , so $\xi_0$ is a constant,
according to Eq.(\ref{ab}), $1/{\xi_1}-1/{\xi'}$ should
be equal to $1/{\xi_0}$ and be constant as the gain varies.
We have checked this prediction and find indeed that the numerical values are
almost the same as the ones predicted theoretically.
From Eq.(\ref{ab}), we can predict the basic features of the
length dependence of $<\ln{T}>$ shown in Fig.~4. When
$1/{\xi_1} > 1/{\xi_0}$,
$<\ln{T}>$ will increase with $L$, and will reach a maximum value when $L$
gets to $L_c$. But if
$1/{\xi_1} < 1/{\xi_0}$, the $<\ln{T}>$ will decrease monotonically,
at first with a slope of $-|1/{\xi_1}-1/{\xi_0}|$
from the origin,
at $ {L_c} $ the curve has a turning point and the slope changes to
$-|1/{\xi_1}+1/{\xi_0}|$. If $1/{\xi_1} \simeq 1/{\xi_0}$,
the curve is almost horizontal for small L and begins to decrease with a
slope
$-1/\xi$ at the critical length. This behavior is exactly shown in Fig.~4.
\par
The critical length $L_c$ is one of the most important parameters of a
random system with gain.
For a random system, one of the most important theories is the Letokhov
theory \cite{Letokhov}. Zhang \cite{Zhang}
generalized the theory and used the no-gain localization length $\xi_0$
to replace the
diffusion coefficient $D$ in the Letokhov theory and obtain that the
critical
length $L_c \simeq \sqrt{ {\xi_0} {l_g} }$, so that we can clearly
see localization effects in the system. But as shown above, there is a
finite critical length in periodic system when the no-gain localization
length $ \xi_0$ goes to infinite, so there must be
other mechanisms for determining the critical length in those systems. We
find that when the localization effect is strong
enough so that the no-gain localization length $ \xi_0 \ll
{(L_c^0)}^2/\xi_1$, then the results of the Letokhov theory are quite good.
But
when the system randomness is weak so that $ \xi_0 $ is larger than
$(L_c^0)^2/{\xi_1}$, then the Letokhov theory results are not correct, and
we have to use other theories, such as
the Lamb theory \cite{LaserPhys}, which is well known is laser physics, to
determine $L_c$. Next we will compare the Letokhov theory
with the Lamb theory, and find the expressions of the critical length in
different cases.
\par
According to the Letokhov theory \cite{Letokhov,Zyuzin2} the field in
the system satisfies:
\begin{eqnarray}
\frac{\partial \phi({\vec r}, t) }{\partial t} =
D{\nabla ^2} {\phi({\vec r}, t) }
+\frac {c{\phi ({\vec r}, t) }}{l_g} \label{Letokhov}
\end{eqnarray}
where $D$ is the diffusion coefficient, $c$ is the speed of the wave.
Considering the relaxation after long time, the solution \cite {Zyuzin2} of
Eq.(\ref{Letokhov}) is :
\begin{eqnarray}
\phi({\vec r}, t)
&\propto& e^{-t[D{(\frac{\pi}{L})}^2 - \frac{c}{l_g}]} \nonumber \\
\frac {\partial \phi({\vec r}, t) } {\partial t} &=&
-D{(\frac{\pi}{L})}^2 \phi({\vec r}, t)
+\frac {c{\phi ({\vec r}, t) }}{l_g}
\end{eqnarray}
When $L={L_c}=\pi \sqrt{ \frac{D {l_g}}{c}} $ , the system is at a
critical point. If $L<L_c$ the field will decay vs time, but if $L>L_c$
then the field in the system will become stronger and stronger with time.
We can clearly see that the physical
meaning of $L_c$ is the balance point of the gain and loss in the system.
When the $L$ is less than $L_c$, the photon escaping rate, which is
determined by $\frac{D {\pi^2}}{L^2}$, is
larger than the photon generating rate, which is determined by
$\frac{c}{l_g}$ of the system, so the
photons generated by the stimulated emission can escape from the system
instantaneously and the system can get to the static state after
a long time. If $L$ is larger than $L_c$, gain
is larger than loss, and
photons will be accumulated in the system \cite{Zyuzin2}.
Based on the Letokhov theory and the weak localization theory
\cite{Anderson,John} results, Zhang \cite{Zhang}
generalized the $L_c$ to be $ L_c \simeq \sqrt{ {\xi_0} {l_g} } $ since
$D = \frac{1}{3}lc$, where $l$ is the mean free path and
$\xi_0 = (2\sim4)l$.
\par
In our models, considering the periodic background, we substitute $\xi_1$
for $l_g$ first. But when the disorder becomes weaker and weaker,
the system become almost periodic, $\xi_0$ goes to infinite,
$L_c$ goes to $L_c^0$ instead to infinite. How one can explain this
behavior of $L_c$? The Lamb theory can give a
theoretical explanation of it.
In the Lamb theory, a phenomenological parameter Q(L),
the quality factor which generally is a function of system length
$L$, is introduced to show the energy loss
rate of the system (also can be thought as the photons loss rate of the
system). In the Lamb theory, the magnitude of the electric field in a
linear medium satisfies the following equation:
\begin{eqnarray} \label{Lamb}
\frac {\partial |E(t)|} {\partial t} = -{\frac {\omega} {2Q(L)}} |E(t)|
+\frac{c}{l_g} |E(t)|
\end{eqnarray}
At the critical condition, the gain term and loss term are equal. We have:
\begin{eqnarray}\label{LambCritical}
\frac {\omega} {2Q({L_c})}=\frac{c}{l_g}
\end{eqnarray}
If we compare this Eq.(\ref{Lamb}) with the solution of Letokhov theory
Eq.(19), we can
find the similarity between them. This similarity is from the same
physical principle, the interplay of loss and gain in the system. From
Eq.(\ref{Lamb}) we can see that the gain term is same as the one given by the
Letokhov theory, the only
difference is from loss term. Generally, ${\frac {\omega} {2Q}}$ is a
function of the system length, e.g. for Fabry-Perot interferometer
${\frac {\omega} {2Q}}\propto {\frac{1}{L}}$ \cite{LaserPhys}. For periodic
system $Q={Q_p}$, we
have ${\frac {\omega} {2Q_p}}\propto {\frac{1}{L}}$ too. From the balance
of gain and loss, we can get that ${L_c^0} \propto {\xi_1}$ in
agreement with our results presented in
section III. This means that in a periodic system the rate of loss is
not infinite, although the no-gain localization length goes to infinite. The
rate is
determined by the ${Q_p}$ of the system and we can get a finite $L_c^0$
correspondingly. From the $L_c^0$ obtained above and the critical condition
given by Eq.(\ref{LambCritical}), we have that the quality factor of the
periodic system is given by:
\begin{eqnarray} \label{Qp}
{Q_p}={\frac {\omega {\xi_1}} {2c{L_c^0}} } L
\end{eqnarray}
which is independent of no-gain localization length $\xi_0$.
For a random system, things are a little more difficult.
The theory of Letokhov doesn't give the detailed information of
localized modes but it gives a localization related quantity $D$, the
diffusion coefficient.
According to the localization theory, $D$ is directly related with
localization
length $\xi_0$, just as Zhang discussed \cite{Zhang}. Based on the
correct results of the
Letokhov theory in the strong localization case, we can assume that when
the disorder is strong enough ${\xi_0}
\ll {{L_c^0}^2}/ {\xi_1} $, the localization effects will dominate
the escape rate of photons of the system (Our numerical results shown in
Fig.~11 support this
assumption). Lamb theory gives
that the Q of a strong random system is determined by the
localization effect. By comparing the corresponding terms in Eq.(19)
and Eq.(\ref{Lamb}), we obtain that:
\begin{eqnarray}\label{Ql}
Q \simeq {Q_l} = \frac {\omega {L^2}} {{\pi^2} D}
= \frac {\alpha\omega {L^2}} {c {\xi_0}}
\end{eqnarray}
where the subscript $l$ is for localized modes, $\alpha$ is a constant
of the order of unity and depends on the ratio of $D$ and $\xi_0$ according
to the localization theory. For both of the models studied here, we find the
$\alpha$ can be chosen to be equal to 0.7. From this we can get
that the critical length
$L_c=\sqrt{{\frac{1}{\alpha}}{\xi_0}{\xi_1}}\simeq\sqrt{{\xi_0}{\xi_1}}$
which is consistent with the Letokhov theory. Eq.(\ref{Ql})
is a very interesting result for laser physics because it is obtained by
the comparison of
the Letokhov and Lamb theories, and it directly gives the
relationship of the quality
factor $Q$ of a random system with the no-gain localization length $\xi_0$
of the system.
\par
In the weak disorder limit ${\xi_0} \gg
{{L_c^0}^2}/{\xi_1} $, $Q \rightarrow Q_p$ and $ L_c \rightarrow L_c^0 $.
In strong disorder limit ${\xi_0} \ll {{L_c^0}^2}/{\xi_1} $, $Q
\rightarrow Q_l$ and $ L_c \rightarrow \sqrt{\xi_0\xi_1} $. For cases where
${\xi_0}$ is
comparable to $ {{L_c^0}^2}/{\xi_1} $, both the effects of periodic
background and randomness will be important to determine the quality
factor of
such a system. Considering the $Q$ as the photon-resistance in the system,
and if we assume that both effects are $independent$ with each other, we
have that the total quality factor of the system to be:
\begin{eqnarray}\label{Q}
Q={Q_p}+{Q_l}={\frac {\omega}{c}} \left(\frac{{\xi_1}L}{L_c^0} +
\frac{\alpha L^2}{\xi_0} \right)
\end{eqnarray}
From the critical condition Eq.(\ref{LambCritical}), we have that:
\begin{eqnarray}\label{Lc}
{L_c}=-\frac{\xi_0 \xi_1}{2{\alpha L_c^0}} +
\sqrt{ {\left(\frac{\xi_0 \xi_1}{2{\alpha L_c^0}}\right)}^2 +
\frac{\xi_0 \xi_1}{\alpha }}
\end{eqnarray}
In Fig.~6a~and~6b, we compare the theoretical predictions given by
Eq.(\ref{Lc}) and by Zhang
\cite{Zhang} with our numerically calculated results for the
classical
many-layered model and the electronic tight-binding model respectively.
Our numerical results shown in Fig.~6a and 6b strongly
support Eq.(\ref{Lc}) to be the correct expression of the critical
length $L_c$ for
both the weak and the strong random limits. In some other cases, the
deviation
can be as large as fifteen percent which is still very good considering
the many approximations that have been introduced in derivation of
Eq.(\ref{Lc}). One explanation for this deviation is that the two
effects that were added in Eq.(\ref{Q}) are not totally $independent$,
because the correlated
scattering of the periodic background will affect both $Q_p$ and $Q_l$.
\section{Probability Distribution of Reflection Coefficient}
\par
Pradhan and Kumar \cite{PR} first obtained the probability
distribution of
the reflection coefficient for a long system with randomness and gain,
which is given by:
\begin{eqnarray}
P(x)=P \left(\frac{R-1}{2q}\right)=\left(\frac{2q}{R-1}\right)}^2
{\exp\left({\frac{-2q}{R-1}}\right) \label{PR}
\end{eqnarray}
where $x=\frac{R-1}{2q}$ and $q={\xi_0}/{\xi_1}$. We numerically
calculated
the $P(x)$ (or $P(R)$) for both models in cases that $q$ changes
drastically. Our numerical features of $P(x)$ ($P(R)$) give some
interesting new results.
According to Eq.(\ref{PR}), the maximum
probability of $P(x)$ (or $P(R)$) should appear at 0.5 (or $ R=q+1 $), and
the distribution has a long tail of large $x$ (or $R$).
For the many-layered model, when the wavelength is near the band center
($\lambda=360nm$), we have that $L_c^0/\xi_1 \simeq 7$ for different
gains from Eq.(\ref{LC0}).
When $q$ at the range of 0.01 to 50, and so ${\xi_0} < {L_c^0}^2/{\xi_1} $,
the $P(x)$ ( $P(R)$) behaves as predicted by Eq.(\ref{PR}).
When $q$ is at the range of 50 to 500, the position of the maximum $P(x)$
($P(R)$)
begins to shift left away from the point 0.5 (or $q+1$ ).
When $q$ increase further, such as to become close to a thousand, the
maximum of $P(R)$ shift to the value of $R_0$,
which is the saturated value of reflection for the periodic system, and
$P(R)$ begin to change its shape
into a delta function and the long tail disappears. This process is
clearly shown in Fig.~7. It is reasonable
to assume that when $q$ is very large, the system is similar to
a periodic system , so the $P(R)$ changes to a delta function which is the
distribution of the reflection coefficient of the periodic system.
\par
For the tight-binding model, when frequency is at band center ($E=0$),
${L_c^0}/{\xi_1}$ is not a constant as $\eta$ changes. It has a range from
5 to 7. When $q$ at
the range of 0.01 to 20(${\xi_0} < {L_c^0}^2/{\xi_1} $), the $P(x)$
( or $P(R)$ ) behaves as predicted by Eq.(\ref{PR}). When
$q$ is at the range of 20-400, then the position of the maximum $P(x)$ (
$P(R)$ ) begins to
shift left away from theoretical value 0.5 ( or $q+1$). When $q$ is
larger than 400, $P(R)$ develops two peaks
, one peak evolves from the original peak, the other one emerges
at $R_0$. When $q$ is even larger, such as thousands, then the original
peak goes down and disappears, and the new peak become higher at $R_0$. At
the same time the long tail disappears, the $P(R)$ also changes to a delta
function at the position of $R_0$. All these changes are shown clearly in
Fig.~8. In Ref.\cite{India}, they also got two peaks
for $P(R)$, but they did not explain that the new peak is due to the
periodic
background of the system and that the delta function is at the position of
$R_0$, the saturated value of periodic system. When q is very small, we
obtain that the $P(x)$ ( or $P(R)$) is almost same as the one predicted by
Eq.(\ref{PR}), quite different from results
of \cite{India}. We think that
this difference is due to the fact that they have not renormalized their
numerical results.
\par
In summary, from our numerical results, we get the general behavior of
$P(x)$ ( or $P(R)$ ) for
both models. When ${\xi_0} < {L_c^0}^2/{\xi_1} $, the $P(x)$ ( or $P(R)$ ) is
same as the theoretically predicted one by Eq.(\ref{PR}). When
${\xi_0}$ is bigger
than ${L_c^0}^2/{\xi_1} $, we must think about the effect of the periodic
background and if ${\xi_0}$ is really very large, the periodic background
will
dominate the behavior of $P(R)$. We also find that at the band edge
wavelength for many-layered model
($\lambda$=470nm) or at the band edge energy for the tight-binding model
$(E=1.8)$,
the results of $P(x)$ ( or $P(R)$ ) are always different from the
predictions given by Eq(\ref{PR}).
We know that at the band edge the effects of coherent scattering will be
very strong and make the long paths of wave propagation more important .
So we think it is this coherent scattering effect which make the $P(x)$ (
or $P(R)$ ) different from the
prediction of Eq(\ref{PR}) which is obtained for homogeneous random systems.
\section{Conclusion}
In conclusion, we have studied the transmission and reflection
coefficients in periodic or
periodically correlated random systems with homogeneous gain.
Theoretically, for periodic systems we predicted the behaviors of
transmission and reflection coefficients, such as the slopes of long
systems and of short systems and critical length by the transmission matrix
method. For random systems, first the Zhang's formula of the
localization length for long systems is checked. We find that only
at the band
edge and with very strong gain and strong disorder, there is
obvious deviation from the theoretical prediction of the localization
length with gain. For short systems, our numerical results show that
our generalization of the formula of absorbing system is correct for
amplifying systems. According to this generalization we can predict the
behaviors of average
value of logarithm of transmission coefficient $<\ln{T}>$ from the
value of
$1/\xi'$, such as: if it is positive then the $<\ln{T}>$ will increase from
origin at slope $1/\xi'$ and generate a peak at $Lc$ and then start to
decrease at slope $-1/\xi$, if it is negative then the
$<\ln{T}>$ will decrease monotonically and has a turning
point at $L_c$ with the slope change from -$1/|\xi'|$ to $-1/\xi$.
\par
To explain the new behavior of the critical length $L_c$ which we got from
our numerical results, we
compare the Letokhov theory with the Lamb theory and give a general
expression for the critical length considering both the effects of
localization and periodic background. With this comparison, we also
construct the relation of the quality factor $Q$ of a random system with the
localization length $\xi$.
\par
We also study the probability distribution of the reflection coefficient
$P(R)$ of random systems with gain. We find some new behaviors of $P(R)$ and
give the
criteria for the range of validity of the different behaviors
and explain it by the influence of the periodic background too.
The study of wave propagation in amplifying random system is a new and
challenging topic. There are still a lot things to be done.
\section{Acknowledgment}
Ames Laboratory is operated for the US Department of Energy by Iowa State
University by Iowa State University under Contract No.W-7405-Eng-82. This
work was supported by the Director for Energy Research office of Basic
Energy Sciences and Advanced Energy Projects and by NATO Grant No.940647.
|
\section{Introduction}
The measurement of production of $D$--mesons in nucleus--nucleus
collisions at CERN SPS energies is a challenging experimental
adventure.
This is due to the short life time and expected low multiplicity
of $D$--mesons which cause the small experimental
signal originating from $D$ decays to be hidden in the large
background composed of combinations of 'non--signal' tracks.
There is, however, increasing interest
in results on charm production in nucleus--nucleus (A+A)
collisions which
motivates various experimental groups to design
upgrades of the existing experiments or to build new
experiments, which should allow to measure open charm
production in A+A collisions \cite{add2,add3,NA50}.
The aim of this paper is to demonstrate that the first measurement
of $D$--meson production may be even possible using the current
set--up of the NA49 experiment \cite{NA49}.
The paper starts with a summary of physics motivations
for charm measurement in A+A collisions at the CERN SPS (Section 2).
The expected multiplicity of $D$--mesons in central Pb+Pb collisions
at 158 A$\cdot$GeV is discussed in Section 3.
An estimate of the
statistical resolution of $D$--meson measurement which can
be achieved using data of the NA49 experiment is given in Section 4.
Summary and conclusions (Section 5) close the paper.
\newpage
\section{Physics Beyond QGP: pQCD vs Thermodynamics }
The main motivation for a broad experimental program
in which nucleus--nucleus collisions at high energies
are studied is the search for the Quark Gluon Plasma
\cite{Co:75,Sh:80}.
An impressive set of data was collected during the last decades
and many unexpected phenomena were discovered
\cite{QM97}.
However, the question whether a transient QGP state is
created in the early stage of the collisions is still
under vivid discussion.
In our opinion this situation is due to the fact
that there is no consensus on which models of
high energy hadronic and nuclear collisions should be
used to interpret experimental results.
The data are in qualitative and quantitative agreement
with the hypothesis of QGP creation in A+A collisions
at the CERN SPS if a statistical model of the early stage
is used for their interpretation \cite{Ga:98}.
This model is, however, considered a non--orthodox
approach to hadronic and nuclear collisions as its basic
assumptions can not be derived from the commonly accepted
theory of strong interaction, QCD.
Its validity is, therefore, more
surprising than the conclusions reached within this model
concerning QGP creation at the CERN SPS.
On the other hand it is difficult to use QCD for the interpretation
of the experimental results. Problems arise because
almost all effects intuitively expected in the case of
a transition to QGP are in the domain of so--called
soft processes in which experimentally testable strict
predictions of QCD are not possible.
Attempts to build phenomenological
QCD--inspired models of soft processes
are not very succesful \cite{Od:98}.
Conclusive interpretation of the data within these models
seems to be impossible as one can not estimate the uncertainties
introduced by the used approximations.
As a possible solution one tries to identify phenomena
sensitive to the early stage in which
so--called hard QCD processes are believed to prevail \cite{Sa:97}.
In this case one hopes to obtain testable QCD predictions
using perturbative methods.
A well known example is the analysis of $J/\psi$
production in A+A collisions performed under the assumption
that the initial production of $c\overline{c}$ pairs is
described by pQCD.
However the limits of the applicability of pQCD are
theoretically not well defined.
Therefore any assumption
based on validity
of pQCD calculations and used to interpret the data
should be tested experimentally.
In the case of the $J/\psi$ example this test can be done
by the measurement of the open charm yield and its comparison to
pQCD predictions.
In connection with the above discussion of QCD--based models
it is interesting to consider a hypothetical model in which,
due to technical problems, strict experimentally testable
predictions can not be obtained.
The model gives approximate predictions, however, the error
made due to the approximations is difficult to estimate.
Experimental results which are in agreement with the predictions
of the model can be treated as a `proof' of its validity.
On the other hand any disagreement between the model and the data
can be interpreted as due to the used approximations.
Therefore the model can not be falsified
\cite{Po:59} before a substantial improvment in its predictive power.
This example illustrates the logical problem
which one may encounter discussing
the validity of QCD--based models.
In the summarized exciting situation
concerning our understanding of strong interactions
the role of the experimental results on open charm production
in A+A collisions is unique.
\begin{itemize}
\item
Data on open charm should allow to test limits of applicability
of the perturbative QCD methods.\\
It is often assumed that the charm quark is
heavy enough for pQCD treatment \cite{Sa:95}.
Based on this hypothesis results on $J/\psi$ production
in A+A collisions are interpreted in the so--called
'suppression' models
\cite{Ge:98, Sa:97},
in which the suppression is calculated relative to the $J/\psi$
yield expected from the pQCD factorization theorem \cite{Co:98}
and data on $J/\psi$ production in p+p interactions.
It was, however, recently shown that the $J/\psi$ multiplicity
is, in good approximation, proportional to pion multiplicity \cite{Ga:98a}.
Thus the $J/\psi$ yield shows an $A$--dependence which is characteristic
of soft QCD processes.
This experimental observation motivates the question whether
pQCD can be used as a model of charm quark production
in hadronic and nuclear collisions.
\item Data on open charm should allow to test the validity
of a non--orthodox statistical model of the early stage
of A+A collisions \cite{Ga:98}. \\
In this model charm quark multiplicity is large enough
to justify the use of the thermodynamical approximation.
Consequently not only the absolute yield of open charm but
also its $A$--dependence are very different for the statistical
and the pQCD--based models.
\end{itemize}
\newpage
\section{ Multiplicity of D--Mesons in Nuclear Collisions}
This section starts with a review of the data on $D^0$ and
$\overline{D}^0$ production in proton--nucleon interactions,
which leads to an estimate of the mean multiplicity of
$D^0 + \overline{D}^0$ mesons, $\langle D^0 + \overline{D}^0 \rangle$,
in nucleon--nucleon, N+N, interactions at 158 A$\cdot$GeV.
This result is further used to predict the multiplicity of
$D^0 + \overline{D}^0$ mesons in central Pb+Pb collisions
at 158 A$\cdot$GeV using four different approaches.
\subsection{$\langle D^0 + \overline{D}^0 \rangle$ in N+N interactions}
There are six measurements of $D$--meson production cross section
at various collision energies
ranging from 200 GeV to 800 GeV, which allow to estimate
the mean multiplicity
of $D^0 + \overline{D}^0$ in p+N interactions
\cite{Be:88, Al:96, Ag:84, Ag:88, Am:88, Le:94}.
In p+p interactions $\sigma(D^0 + \overline{D}^0)$ is measured at
360 GeV \cite{Ag:84}, 400 GeV \cite{Ag:88} and at 800 GeV \cite{Am:88}.
At 200 GeV \cite{Be:88}, 250 GeV \cite{Al:96} and 800 GeV \cite{Le:94}
the cross section in p+N interactions is estimated based on
measured cross section for p+A interactions assuming
that $\sigma(D^0 + \overline{D}^0)$ is proportional to $A$.
We note here that this assumption is not fully justified by the
experimental results \cite{Ga:98a} and may lead to an additional
systematic error in estimating $\sigma(D^0 + \overline{D}^0)$
in p+N interactions.
In addition a symmetry of the $x_F$ distribution of $D^0 + \overline{D}^0$
mesons
with respect to $x_F = 0$ is assumed for the calculation of the integrated
cross section using measurements at $x_F > 0$.
In the case of the measurement at 200 GeV \cite{Be:88} only results
for the sum of
all
$D$--mesons are published; in this case we assume that 50\% of
them are $D^0 + \overline{D}^0$ mesons.
The mean multiplicity of $D^0 + \overline{D}^0$ mesons in p+p
and p+N interactions is calculated as a ratio
$\sigma(D^0 + \overline{D}^0)/\sigma$, where $\sigma$
is the inelastic cross section for p+p interaction at the
corresponding energy.
The values of $\langle D^0 + \overline{D}^0 \rangle_{NN}$ are given in Table 1
and plotted as a function of $\sqrt{s}$ in Fig. 1.
The results at $\sqrt{s} = 19.4$ GeV and $\sqrt{s} = 21.7$ GeV
differ by a factor of about 8, which suggests a possible large
systematic uncertainty of the measurements at low energy not
accounted for in the quoted errors.
Taking this into account we estimate the mean multiplicity
of $D^0 + \overline{D}^0$ at $\sqrt{s} = 17.3$ GeV ($E_{LAB} = 158$ GeV)
in nucleon--nucleon interactions
by an arithmetic average of the measurements at $\sqrt{s} = 19.4$ GeV
and $\sqrt{s} = 21.7$ GeV, which gives:
\begin{equation}
\langle D^0 + \overline{D}^0 \rangle_{NN} = 2.0 \cdot 10^{-4}.
\end{equation}
A large systematic error up to several times the estimated value
should be kept in mind.
\subsection{$\langle D^0 + \overline{D}^0 \rangle$ in Central Pb+Pb Collisions
at 158 A$\cdot$GeV}
We discuss here four approaches which allow to estimate the mean
multiplicity of $D^0 + \overline{D}^0$ in central Pb+Pb collisions,
$\langle D^0 + \overline{D}^0 \rangle_{PbPb}$, at 158 A$\cdot$GeV.
It is assumed that the central collisions are selected by
accepting only events with a low number of spectator nucleons
from the projectile nucleus and that typical trigger conditions
of the NA49 experiment are used \cite{NA49}.
This results in a mean number of participant nucleons,
$\langle N_P \rangle$,
of
about 350 \cite{PRL}.
\vspace{0.2cm}
\noindent
{\bf A}. It was recently observed \cite{Ga:98a}
that the mean multiplicity of $J/\psi$
mesons increases proportinally to the mean multiplicity of
negatively charged hadrons (more than 90\% $\pi^-$ mesons)
when p+p, p+A and A+A collisions are considered (see Fig. 2).
It may be therefore assumed that a similar dependence on the
size of the colliding objects is also valid for $D$--mesons.
Following this assumption a mean multiplicity of
$D^0 + \overline{D}^0$ mesons can be calculated as:
\begin{equation}
\langle D^0 + \overline{D}^0 \rangle_{PbPb} =
\langle D^0 + \overline{D}^0 \rangle_{NN} \cdot
\frac {\langle h^- \rangle_{PbPb}}
{\langle h^- \rangle_{NN}} = 2 \cdot 10^{-4} \cdot
\frac {700} {3.1} \approx 4.5 \cdot 10^{-2},
\end{equation}
where the values of $\langle h^- \rangle_{PbPb}$ and $\langle h^- \rangle_{NN}$
are taken from Ref. \cite{PRL} and Ref. \cite{Ga:95}, respectively.
\vspace{0.2cm}
\noindent
{\bf B}. It is usually assumed that due to the large mass of the charm quark,
production of charm can be calculated using perturbative QCD methods
\cite{Sa:95}.
This assumption leads to the expectation that the cross section for charm
production increases as $A^2$ for all inelastic A+A collisions and
that the mean multiplicity of open charm increases as
$\langle N_P \rangle^{4/3}$ for central A+A collisions.
Thus for central Pb+Pb collisions we get:
\begin{equation}
\langle D^0 + \overline{D}^0 \rangle_{PbPb} =
\langle D^0 + \overline{D}^0 \rangle_{NN} \cdot (\langle N_P \rangle/2)^{4/3}
\approx 2 \cdot 10^{-1}.
\end{equation}
This estimate agrees with a previously published prediction
for charm production based on pQCD inspired models \cite{Br:98}.
\vspace{0.2cm}
\noindent
{\bf C}. The NA50 Collaboration found recently \cite{Sc:98}
that a model based on pQCD can not describe the dimuon invariant mass
spectrum between the $\phi$ and $J/\psi$ peaks in central Pb+Pb
collisions.
The spectrum, however, can be reproduced when the contribution from
open charm decays is scaled up by a factor of about 3.
Based on this analysis one may expect that:
\begin{equation}
\langle D^0 + \overline{D}^0 \rangle_{PbPb}
\approx 6 \cdot 10^{-1},
\end{equation}
i.e. it is equal to the multiplicity calculated in {\bf B}
multiplied by a factor of 3.
\vspace{0.2cm}
\noindent
{\bf D}. The production of entropy and strangeness in A+A collisions
at the CERN SPS can be described by a statistical model which assumes
the creation of a Quark Gluon Plasma in the early stage of the
collision \cite{Ga:98}.
This fact and the observation that the $J/\psi$ yield is
proportional to the pion yield triggered the hypothesis
that also charm production can be described by the
same statistical approach.
It was calculated, within this model \cite{Ga:98},
that the mean number of $c$ and $\overline{c}$
quarks produced in central Pb+Pb collision at 158 A$\cdot$GeV
is about 17.
Based on p+p data \cite{Ag:88}
we assume here that about one third of them hadronize
as $D^0$ and $\overline{D}^0$ mesons which gives:
\begin{equation}
\langle D^0 + \overline{D}^0 \rangle_{PbPb}
\approx 6.
\end{equation}
\vspace{0.5cm}
\noindent
We summarize this section by observing that predictions
of the mean multiplicity of $D^0 + \overline{D}^0$ mesons
in central Pb+Pb collisions at 158 A$\cdot$GeV
range from 4.5$\cdot$10$^{-2}$ to about 6 showing more than
a factor 100 difference between minimum and maximum values.
\newpage
\section{$D$--Meson Measurement in NA49}
We consider here the possibility of a measurement of $D^0$
and $\overline{D}^0$ production in A+A collisions
at 158 A$\cdot$GeV using the current set--up of the NA49
experiment at the CERN SPS \cite{NA49}.
We start from an introductory discussion on
the means to distinguish $D^0$ decays
from combinatorial background and on the calculation of
the statistical error on $D^0$ multiplicity.
We continue with a brief summary of the relevant properties
of the NA49 experiment and finally we present
results of a simulation which yields an estimate of
the statistical resolution
of $D^0 + \overline{D}^0$ measurement
in central Pb+Pb collisions
using the NA49 data.
Our analysis is based on the study of the two body
decay channels of $D^0$ and $\overline{D}^0$ mesons:
\begin{equation}
D^0 \rightarrow K^- + \pi^+ {\rm ~~~~~and~~~~~}
\overline{D}^0 \rightarrow K^+ + \pi^-
\end{equation}
for which the branching ratio is measured to be
(3.85 $\pm$ 0.09)\% \cite{PDG}.
The mass and the proper life time
of $D^0$ and $\overline{D}^0$ mesons are
$m_{D^0}$ = (1864.6 $\pm$ 0.5) MeV/c$^2$ and
$c \tau$ = 0.01244 cm, respectively \cite{PDG}.
\subsection{Introductory Remarks}
Let us consider central Pb+Pb collision at CERN SPS energy
in which among thousands of produced particles are also $D^0$
mesons.
The $D^0$ mesons decay
after a typical flight distance of about 0.1 cm
($\approx \gamma \cdot c \tau$) from the
collision point.
There are many decay channels possible, but for the reasons
discussed at the end of this paper we consider only decays of
a single type, $D^0 \rightarrow K^- + \pi^+$,
which happen only in 3.85 \% of all cases of $D^0$ decays.
For simplicity of the initial discussion we assume
that an ideal detector is placed around the collision
point,
i.e. for all charged particles electric charge, mass and momentum vector
at a given reference plane are measured.
For all ($K^-$, $\pi^+$) pairs originating from $D^0$ decays
(signal pairs) the following conditions have to be
approximately (within experimental resolution) fulfilled:
\begin{enumerate}
\item
the trajectories of $K^-$ and $\pi^+$ intersect in a point
(decay point) which is different from the Pb+Pb collision point,
\item
the distribution of the $D^0$ life time in its own c.m.s. frame is an
exponential distribution with a mean value equal to
$c \tau = 0.0124$ cm,
\item
energy and momentum conservation laws are fulfilled at the decay
point when $D^0$ decay hypothesis is assumed,
\item
the angular distribution of the decay products in the $D^0$ c.m.s. is
isotropic.
\end{enumerate}
Conditions 1--4 can be fulfilled
(within experimental resolution) by a set of ($K^-$, $\pi^+$) pairs
which do not originate from the same $D^0$ decay (background pairs)
only by chance.
The probability of this happening decreases with increasing experimental
resolution.
As a useful example we consider the condition 3 of energy momentum conservation
at the decay vertex.
For practical reasons one often quantifies deviations from
energy momentum conservation, assuming the $D^0$ decay
hypothesis,
by calculating
the invariant mass, $M(K^-, \pi^+)$, for a pair
of negatively and positively charged particles and
comparing it to the known mass of the $D^0$ meson.
Almost all signal pairs are expected to be distributed in a narrow
interval $\Delta M$ centered around $m_{D^0}$.
The $M$ distribution for background pairs
is significantly broader and it has no characteristic peak
structure at $m_{D^0}$.
The smearing of the $M$ distribution for signal pairs results
from the finite resolution of momentum measurement.
With increasing resolution the size of the interval $\Delta M$
in which almost all signal pairs are included decreases
consequently resulting in lower number of accepted background pairs.
In the high resolution limit almost only signal pairs
(and almost no background pairs)
are selected when taking pairs from $\Delta M$.
In this case the total number of signal pairs, $N_S$, in $N_{EV}$
events can be approximated by the measured number of pairs in the
$\Delta M$ interval, $N_{\Delta M}$.
An estimate of the mean multiplicity of signal pairs
is therefore given by
\begin{equation}
\langle N^E_S \rangle = \frac { N_{\Delta M} } { N_{EV} }
\end{equation}
and its statistical error can be calculated as
\begin{equation}
\sigma (\langle N^E_S \rangle) = \frac
{ \sqrt{\langle N_{\Delta M} \rangle} } { \sqrt{N_{EV}} } \approx
\frac
{ \sqrt{\langle N^E_S \rangle} } { \sqrt{N_{EV}} }
\end{equation}
assuming a Poissonian distribution of pair
multiplicity in the interval $\Delta M$.
In the limit of poor resolution the number of background pairs, $N_B$,
in the interval $\Delta M$ is much larger than the number of
signal pairs.
Therefore
in order to estimate the number of signal pairs, $N^E_S$, an
estimate of the number of background pairs, $N^E_B$ is needed:
\begin{equation}
N^E_S = N_{\Delta M} - N^E_B
\end{equation}
The estimate of $N^E_B$ is model dependent and therefore
it has a systematic error which however will not be discussed here.
Various models can be used.
As an example let us mention a 'mixed event' model of the background
in which background pairs are constructed by selecting
particles from different events.
Independent of the model used it is reasonable to assume
that the statistical error of $N^E_B$ can be made much smaller
than the statistical error of $N_{\Delta M}$.
Thus the statistical error of $\langle N^E_S \rangle$ can
be calculated as
\begin{equation}
\sigma(\langle N^E_S \rangle) = \frac { \sqrt{\langle N_{\Delta M} \rangle} }
{ \sqrt{N_{EV} } } \approx
\frac { \sqrt{\langle N^E_B \rangle} } { \sqrt{N_{EV} } },
\end{equation}
assuming Poissonian distribution of pair multiplicity in the
interval $\Delta M$.
We observe that in the limit of high resolution the statistical
error of the signal is almost fully defined by the number of
signal pairs, whereas in the limit of poor resolution it is defined
by the number of background pairs.
One can define the statistical significance of the measurement
as the ratio $\langle N^E_S \rangle/\sigma(\langle N^E_S \rangle$
which in the case of the poor resolution limit is given by
\begin{equation}
\frac {\langle N^E_S \rangle} {\sigma(\langle N^E_S \rangle)} =
\frac {\langle N^E_S \rangle} {\sqrt{\langle N^E_B \rangle}}
\cdot \sqrt{ N_{EV} }.
\end{equation}
Thus in order to maximize the statistical significance of the result
one should select the acceptance (in the example above
the size of the $\Delta M$ interval) such that the ratio
$\langle N^E_S \rangle/\sqrt{\langle N^E_B \rangle}$
reaches a maximum.
\subsection{The NA49 Experiment}
The NA49 experiment \cite{NA49} at the CERN SPS was designed
and constructed to search for signals of the Quark Gluon Plasma
created at the early stage of nucleus--nucleus collisions.
Basic detectors of the NA49 set--up are four Time Projection
Chambers (TPC's), which allow for a precise tracking
of charged particles.
Two medium size TPC's of 3 m$^3$ gas volume each are located
inside of two magnets with up to 1.5 T field strength each.
Two large size TPC's of 20 m$^3$ gas volume each are positioned
downstream of the magnets for high precision energy loss measurement
and acceptance coverage in the forward direction.
Overall the TPC's acceptance coverage amounts to up to 80\%
of all produced charged particles (this
number depends on the reaction studied and magnetic field --
target configuration).
The TPC's measure up to 234 space points on tracks of up
to 13 m length.
This allows a precise determination of sign of electric
charge, momentum vector ($\sigma(p)/p^2 \approx 10^{-4}$ (GeV/c)$^{-1}$)
and energy loss ($ \approx 4$\% relative resolution)
for all accepted particles.
These measurements yield information on particle mass
which results in large acceptance (limited resolution)
particle identification.
Four Time--of--Flight walls
($\approx $ 60 ps resolution) complement the particle identification
capabilities of the NA49 detector.
A typical resolution in reconstruction of the distance
between collision point and secondary vertex of a two body decay
near the target
is of the order of 1 cm and is mainly due to the long extrapolation length
needed from the first TPC detector.
\subsection{Statistical Error on
$\langle D^0 + \overline{D}^0 \rangle_{PbPb}$
in NA49}
In order to estimate the statistical resolution of
a $\langle D^0 + \overline{D}^0 \rangle_{PbPb}$ measurement in NA49
we performed a simulation of signal and background pairs
as expected in this experiment for central
Pb+Pb collisions at 158 A$\cdot$GeV.
In the simulation the standard geometry and magnetic field of
NA49 are assumed \cite{NA49}.
We define the geometrical accpetance of the NA49 TPC's
by the requirement that the particle trajectories cross more than 20
TPC padrows, at least one of which has to be in a TPC located
inside the magnetic field.
A parametrization of the momentum resolution as measured by NA49
is included in the simulation.
Concerning particle identification we consider two extreme
cases.
In the first one we assume that no information on particle mass
is available, whereas ideal particle identification is
assumed in the second case.
For the background calculation spectra of charged hadrons
produced in central Pb+Pb collisions at 158 A$\cdot$GeV
as measured by NA49 \cite{PRL} are parametrized in rapidity ($y$)
and transverse momentum ($p_T$).
The signal simulation is done assuming a Gaussian ($\sigma = 0.6$)
rapidity distribution and an exponential spectrum in
transverse mass ($T = 300$ MeV) for both $D^0$ and
$\overline{D}^0$ mesons.
The total multiplicity of $D^0$ and $\overline{D}^0$ mesons of 6
as predicted by
the statistical model (see point 3.2.{\bf D}) is assumed (if needed).
In Fig. 3 we show the $y-p_T$ distribution of $D^0$ mesons
for which both decay products are in the geometrical acceptance of
the NA49 TPC's.
Before further presentation of results of the simulation we note that
conditions 1 and 2 (see Section 4.1),
which require an accurate measurement of the decay vertex,
can not be used in the exisiting NA49 set--up
to distinguish signal from background pairs.
This is due to the poor resolution in the
reconstruction of the secondary vertex
from two body decays of the order of 1 cm, which is
much larger than the typical flight path of $D^0$ and $\overline{D}^0$
of about 1 mm.
Therefore background rejection has to be done using conditions
3, energy--momentum conservation, and 4,
isotropy of the decay, only.
We quantify these conditions by studing $M(K^-, \pi^+)$
(or $M(K^+, \pi^-)$) distribution and cos$\Theta$
distribution, where $\Theta$ is the angle between
$D$ and $\pi$ meson directions calculated in $D$--meson c.m. system.
The results of the simulation shown below are obtained for the $D^0$ decay,
without using information on particle
mass.
An improvement of the statistical resolution of charm
measurement due to the addition of $\overline{D}^0$ decays and
particle mass information is considered at the end of this section.
In Fig. 4 we show an invariant mass distribution $M(K^-,\pi^+)$
for background pairs obtained using the geometrical acceptance of
NA49.
It is seen that the $D^0$ meson mass is located in the region of
the monotonically decreasing tail of the distribution,
well beyond the position
of its maximum.
The $M(K^-,\pi^+)$ distributions for signal and background pairs
in the interval of $M$ around $m_{D^0}$ are shown in Figs. 5a and
5c, respectively.
For pairs from the interval $\Delta M$ = 4 MeV/c$^2$
around $m_{D^0}$ we plot also the cos$\Theta$ distributions
in Figs. 5b and 5d.
We observe that in the case of $D^0$
measurement by NA49 we are in the poor resolution limit
i.e. the statistical resolution of signal measurement
is determined by the number of background pairs and
is independent of the signal multiplicity.
It is also clear that rejection of pairs with high
cos$\Theta$ values should result in an improvement
of statistical significance of the signal measurement.
We estimate that the maximum significance can be achieved by
accepting pairs with cos$\Theta < 0.7$ and
inside an interval of $\Delta M = $ 4 MeV/c$^2$ (background cuts).
The mean multiplicity of background pairs for this
selection is
\begin{equation}
\langle N_B \rangle_{NOID} \approx 80,
\end{equation}
where the subscript $_{NOID}$ is used to underline that the number is
obtained without using information on particle identification.
This yields an estimate of the statistical
resolution of $\langle N^E_S \rangle$
\begin{equation}
\sigma_{NOID}(\langle N^E_S \rangle) =
\frac { \sqrt{\langle N_B \rangle}_{NOID} } { \sqrt{N_{EV} } } \approx
\frac { 9 } { \sqrt{N_{EV} } }.
\end{equation}
The resulting correction factors
needed to obtain $\langle D^0 \rangle$ from the estimated signal multiplicity
$\langle N^E_S \rangle$ in the acceptance are:
\begin{itemize}
\item
$w_{BR}$ = 26.0, for the branching ratio,
\item
$w_{GA}$ = 2.4, for the geometrical accepatnce,
\item
$w_{BR}$ = 2.3, for the background cuts.
\end{itemize}
Therefore the statistical resolution of $\langle D^0 \rangle$
can be estimated as:
\begin{equation}
\sigma_{NOID}(\langle D^0 \rangle) = w_{BR} \cdot w_{GA} \cdot w_{BC} \cdot
\frac { \sqrt{\langle N_B \rangle_{NOID}} } { \sqrt{N_{EV} } } \approx
\frac { 1300 } { \sqrt{N_{EV} } }.
\end{equation}
In the case of ideal particle identification background multiplicity
can be reduced by about a factor of 10, which results in
$\sigma_{ID}(\langle D^0 \rangle) \approx 400/\sqrt{N_{EV} }$.
Therefore
for $10^6$ central Pb+Pb collisions at 158 A$\cdot$GeV
$\sigma(\langle D^0 \rangle) \approx$ 0.4--1.3 in comparison to
$\langle D^0 \rangle \approx 3$ expected in the case of
the statistical model.
Statistical significance of measurement of
$\langle D^0 + \overline{D}^0 \rangle$ is approximately by a factor
of $\sqrt{2}$ better than significance of separate measurements for
$D^0$ or $\overline{D}^0$.
In Fig. 6 we plot $\sigma_{NOID}(\langle D^0 + \overline{D}^0 \rangle)$
and $\sigma_{ID}(\langle D^0 + \overline{D}^0 \rangle)$ as a
function of the number of central Pb+Pb collisions at 158 A$\cdot$GeV
registered by NA49.
The different predictions concerning
$\langle D^0 + \overline{D}^0 \rangle$ discussed in Section 3
are also indicated in Fig. 6 for comparison.
Finally in Fig. 7 we show $M(K^-, \pi^+)$ distributions obtained
using the cos$\Theta$ cut for signal and background pairs for
100 central Pb+Pb collisions.
In this case the number of generated $D^0$ decays was scaled up by a
factor of 100 and therefore the statistical significance of the
observed signal peak corresponds to the significance
expected for $10^6$ events.
\subsection{Discussion}
The experiment NA49 registered up to now about 10$^6$
central Pb+Pb collisions at 158 A$\cdot$GeV.
As follows from the results of simulation presented in Fig. 5
already the analysis of this data should yield significant results
concerning open charm production in Pb+Pb collisions.
This encouraging conclusion is reached mainly due to
three factors:
\begin{itemize}
\item
there are expectations of a significant enhancement of
open charm production in A+A collisions,
\item
a significant fraction of produced $D$--mesons is covered
by the large geometrical acceptance of NA49,
\item
the good momentum resolution of NA49 allows for a significant reduction of
background even without reconstructing the $D$--meson decay vertex.
\end{itemize}
The results presented in this paper are obtained by a simple
analysis of simulated data.
This scheme was selected in order to
underline the main concepts and to build intuition concerning
basic ingredients of the problem of $D$--meson measurement
in A+A collisions.
The use of more sophisticated statistical methods
of data analysis (see Ref. \cite{Ga:92}), which e.g. include
explicitly the statistical errors of the measured momentum vector
by using a kinematical fit, may lead to an improvement of
the achieved resolution.
We studied also the statistical resolution of open charm measurement
using three and four body decays of neutral and charged $D$--mesons.
We found that due to high combinatorial background the
statistical resolution is
much lower than in the case of the two body decay channel.
\newpage
\section{Summary and Conclusions}
The main results presented in this paper can be summarized as follows.
\begin{itemize}
\item
The measurement of open charm production in A+A collisions
gives a unique possibility to test predictions of thermodynamical
and pQCD--based approaches in the region where both are formally
applicable.
\item
Various approaches used to estimate the $D$--meson multiplicity
in central Pb+Pb collisions at 158 A$\cdot$GeV give predictions
which differ by more than a factor 100.
\item
Experimental data already registered by NA49 should allow to obtain
a significant result on open charm production in Pb+Pb collisions.
The analysis can yield the first observation of open charm
signal or it will lead to an estimate of the upper limit
of open charm multiplicity which should significantly narrow the range
of allowed models.
\end{itemize}
It is obvious that a significant increase of statistics of A+A collisions
at maximum SPS energy registered by NA49 is required for
continuation of open charm program.
This can be achieved during the already scheduled heavy ion run
at the CERN SPS
in the year 2000 and the possible runs beyond.
It is also clear that for a precise measurement of open charm
production an upgrade of the NA49 set--up by a vertex detector, which
allows for an accurate reconstruction of $D$--meson decay vertices,
is needed.
\vspace{0.5cm}
{\bf Acknowledgements}
We thank H. G. Fischer, L. L. Frankfurt, M. I. Gorenstein,
P. Seyboth, R. Stock, H. Str\"obele and G. Roland for
numerous discussions on the subjects coverd in this paper
and comments to the manuscript.
\newpage
|
\section{Introduction}
Consider the sorts of configurations that can be constructed out of a
sequence of line segments, glued end to end to end to form an embedded
loop in $ \Field{R}^{3}. $ The line segments might represent bonds between
atoms in a polymer, segments in the base-pair sequence of a circular
DNA macromolecule, or simply thin wooden sticks attached with flexible
rubber joints. Thus, a spatial polygon of this kind serves as a
mathematical model for some object which is physically knotted yet
retains some of the rigidity inherited from the materials from which
it is built.
It is a classical result of three-dimensional topology that knotted
loops made out of flexible string can always be approximated by
polygonal loops consisting of many thin, rigid segments. Furthermore,
any deformation performed on the string can always be approximated by
a deformation of the polygon, as long as the number of edges is
allowed to increase. However if we insist that the number of edges
remain constant, then we clearly restrict the types of knots that we
can construct. For instance, if we use five or fewer edges, every
loop we build is topologically unknotted; on the other hand we can
build a trefoil or a figure-eight knot if we use six or seven edges,
respectively. See Figure ~\ref{fig:polygons}.
What is not clear is whether we can always mimic a topological
deformation by a deformation of polygons when we place restrictions on
the number of edges. For instance, it is unknown whether we can build
a really complicated polygon which, if it were made out of flexible
string, could be topologically deformed into a round unknot but, if it
were built out of rigid sticks with flexible joints, could not be
flattened out into a planar polygon. In other words, it is an open
question whether there exist topological
unknots which are geometrically knotted.
\begin{figure}[t]
\insertfig{polygons.eps}{1.25in}
\caption{A hexagonal trefoil knot and a heptagonal figure-eight knot.}
\label{fig:polygons}
\end{figure}
As it turns out, it is not always possible to find a geometric isotopy
({\em i.e.} one which keeps the number of edges fixed) between two
polygonal configurations which are topologically equivalent. In fact,
even the case of hexagonal trefoils is nontrivial, as there are
distinct geometric isotopy types, or {\em isotopes}, of this knot. As
a consequence, familiar properties such as reversibility behave
differently when dealing with geometric knots.
One formulation due to Dick Randell \cite{Randell:conform1,
Randell:conform2} is obtained by observing the correspondence between
$ n $-sided polygonal loops in Euclidean three-space and points in $
\Field{R}^{3n}. $ Suppose that $ P $ is an $ n $-sided polygon in $ \Field{R}^{3},
$ together with a choice of a ``first vertex'' $ v_{1} $ and an
orientation. By listing the coordinates of each vertex in sequence,
we obtain a point $ (x_{1}, y_{1}, z_{1}, x_{2}, y_{2}, z_{2}, \ldots,
x_{n}, y_{n}, z_{n}) \in \Field{R}^{3n} $ which we associate with $ P =
\langle v_{1}, v_{2}, \ldots, v_{n} \rangle. $ As in the theory of
Vassiliev invariants, let the {\em discriminant} $ \Sigma^{(n)} $ be
the set of all points in $ \Field{R}^{3n} $ which correspond to polygons with
self-intersections. If $ n > 3, $ this discriminant is the union of $
\frac{1}{2} n (n - 3) $ pieces, each of which corresponds to the set
of polygons with an intersecting pair of non-adjacent edges. For
instance, the subset in $ \Sigma^{(n)} $ consisting of polygons for
which the edges $ v_{1}v_{2} $ and $ v_{3}v_{4} $ intersect can be
described as the collection of polygons for which:
\begin{enumerate} \renewcommand{\labelenumi}{(\roman{enumi})}
\item the vertices $ v_{1}, v_{2}, v_{3}, $ and $ v_{4} $ are coplanar,
\item the line determined by $ v_{1} $ and $ v_{2} $ separates $ v_{3} $
from $ v_{4} , $ and
\item the line determined by $ v_{3} $ and $ v_{4} $ separates $ v_{1} $
from $ v_{2} . $
\end{enumerate}
Note that this set corresponds to the closure of the locus in $
\Field{R}^{3n} $ of the system
\begin{gather*}
(v_{2}-v_{1}) \times (v_{3}-v_{1}) \cdot (v_{4}-v_{1}) = 0 , \notag \\
(v_{2}-v_{1}) \times (v_{3}-v_{1}) \cdot (v_{2}-v_{1}) \times
(v_{4}-v_{1}) < 0 , \\
(v_{4}-v_{3}) \times (v_{1}-v_{3}) \cdot (v_{4}-v_{3}) \times
(v_{2}-v_{3}) < 0 . \notag
\end{gather*}
Therefore, each of these pieces is the closure of a codimension one
cubic semi-algebraic variety, {\em i.e.} a hypersurface with boundary.
We define the space of geometric knots to be the complement of this
discriminant, $ \mathfrak{Geo}^{(n)} = \Field{R}^{3n} - \Sigma^{(n)}. $ Therefore $
\mathfrak{Geo}^{(n)} $ is a dense open submanifold of $ \Field{R}^{3n} . $ In this
space, points correspond to embedded polygons or {\em geometric
knots}, paths correspond to {\em geometric isotopies}, and
path-components correspond to {\em geometric knot types}.
By a theorem of Whitney \cite{Whitney:varieties}, for any given $ n $
there are only finitely many path-components in $ \mathfrak{Geo}^{(n)} .$ It is
also a well-known ``folk theorem,'' due perhaps to Kuiper, that the
spaces $ \mathfrak{Geo}^{(3)}, \mathfrak{Geo}^{(4)}, $ and $ \mathfrak{Geo}^{(5)} $ are connected.
In \cite{Calvo:thesis, Calvo:hexagons}, I showed that the spaces $
\mathfrak{Geo}^{(6)} $ and $ \mathfrak{Geo}^{(7)} $ have five components each. Contrast
this with the fact that only three topological knot types are
represented in $ \mathfrak{Geo}^{(6)} , $ and that only four topological knot
types are present in $ \mathfrak{Geo}^{(7)} .$ When $ n > 8, $ the exact number
of path-components remains unknown. In fact, even the number of
topological knot types represented in the different components of $
\mathfrak{Geo}^{(n)} $ is known only when $ n < 9 .$ The following theorem
summarizes the current status of the classification of geometric knots
with a small number of edges.
\begin{Thm} \label{thm:classification}
(Calvo \cite{Calvo:thesis, Calvo:hexagons})
\begin{enumerate} \renewcommand{\labelenumi}{(\roman{enumi})}
\item The spaces $ \mathfrak{Geo}^{(3)}, \mathfrak{Geo}^{(4)}, $ and $ \mathfrak{Geo}^{(5)} $ are
path-connected and consist only of unknots.
\item The space $ \mathfrak{Geo}^{(6)} $ of hexagonal knots contains five
path-components. These consist of a single component of unknots, two
components of right-handed trefoils, and two components of left-handed
trefoils.
\item The space $ \mathfrak{Geo}^{(7)} $ of heptagonal knots contains five
path-components. These consist of a single component of unknots and
of each type of trefoil knot, and two components of figure-eight knots.
\item The space $ \mathfrak{Geo}^{(8)} $ of octagonal knots contains at least
twenty path-components. However, the only knots represented in this
space are the unknot, the trefoil knot, the figure-eight knot, every
five and six crossing prime knot $ (5_1, 5_2, 6_1, 6_2, $ and $ 6_3) ,
$ the square and granny knots $ (3_1 \pm 3_1) , $ the (3, 4)-torus
knot $ (8_{19}) , $ and the knot $ 8_{20} . $
\end{enumerate} \end{Thm}
It is important to note that although the deformations obtained as
paths in $ \mathfrak{Geo}^{(n)} $ preserve the polygonal structure of the knot
in question, in general they will not preserve edge length. Let $ f:
\mathfrak{Geo}^{(n)} \rightarrow \Field{R}^{n} $ be the map taking $ P = \langle v_{1},
v_{2}, \ldots, v_{n} \rangle $ to the $ n $-tuple \[
(\norm{v_{1}-v_{2}}, \norm{v_{2}-v_{3}}, \ldots, \norm{v_{n-1}-v_{n}},
\norm{v_{n}-v_{1}}) . \] Then points in the preimage $ \mathfrak{Equ}^{(n)} =
f^{-1}(1, 1, \ldots, 1) $ correspond to {\em equilateral knots} with
unit length edges. Since the point $ (1, 1, \ldots, 1) $ is a regular
value for $ f, $ the space $ \mathfrak{Equ}^{(n)} $ is a $ 2n $-dimensional
submanifold (in fact, a codimension $ n $ quadric hypersurface)
intersecting a number of the components of $ \mathfrak{Geo}^{(n)} , $ some
perhaps more than once. Paths in this submanifold correspond to
geometric isotopies which do preserve edge length, so the
path-components of this space offer yet another notion of knottedness.
In his original papers on molecular conformation spaces
\cite{Randell:conform1, Randell:conform2}, Randell shows that if $ n
\le 5 $ then $ \mathfrak{Equ}^{(n)} $ is connected. The case when $ n = 6 $ had
virtually remained untouched for ten years, except for work by Kenneth
Millett and Rosa Orellana showing that $ \mathfrak{Equ}^{(6)} $ contains a
single component of topological unknots.~\!\!\footnote{\, Their
unpublished result is mentioned in Proposition 1.2 of
\cite{Millett:random.knot}.} By focusing attention to a special case
of singular ``almost knotted'' hexagons, \cite{Calvo:hexagons} shows
that two hexagons are equilaterally equivalent exactly when they are
geometrically equivalent. Thus $ \mathfrak{Equ}^{(6)} $ intersects each
component of $ \mathfrak{Geo}^{(6)} $ exactly once. Furthermore,
\cite{Calvo:hexagons} shows that this correspondence of
path-components is not uninteresting, as the inclusion $ \mathfrak{Equ}^{(6)}
\hookrightarrow \mathfrak{Geo}^{(6)} $ has a nontrivial kernel at the level of
fundamental group. In fact, if $ \mathcal{T} $ is a component of
trefoils in $ \mathfrak{Geo}^{(6)} ,$ then $ \pi_{1}(\mathcal{T}) = \Field{Z}_{2} $
while $ \pi_{1}(\mathcal{T} \cap \mathfrak{Equ}^{(6)}) $ contains an infinite
cyclic subgroup.
This paper presents two key ingredients from \cite{Calvo:thesis,
Calvo:hexagons} used to obtain Theorem~\ref{thm:classification}. In
Section~\ref{sec:stratify}, we discuss a method of decomposing $
\mathfrak{Geo}^{(n)} $ into three-dimensional fibres or ``strata.'' This method
proved particularly useful in the analysis of $ \mathfrak{Geo}^{(6)} $ and $
\mathfrak{Geo}^{(7)} .$ Then Section ~\ref{sec:project} describes an upper bound
on the minimal crossing number of the knot realized by an $ n $-sided
polygon. This bound, which is obtained by looking at a particular
projection of polygon into a sphere, improves the one previously known
by a linear term and provides enough control to classify the
topological knot types present in $ \mathfrak{Geo}^{(8)} . $
\section{A Stratification of Geometric Knot Spaces} \label{sec:stratify}
Consider the map $ g $ with domain $ \mathfrak{Geo}^{(n)} $ which ``forgets''
the last vertex of a polygon, mapping
\[ P = \langle v_{1}, v_{2}, v_{3}, \ldots, v_{n-1}, v_{n} \rangle
\mapsto g(P) = \langle v_{1}, v_{2}, v_{3}, \ldots, v_{n-1} \rangle .\]
Notice that a generic polygon in $ \mathfrak{Geo}^{(n)} $ will map to an
embedded polygon in $ \mathfrak{Geo}^{(n-1)} ; $ the only polygons which do not
are the ones for which some part of the linkage $ v_{1} v_{2} \ldots
v_{n-1} $ passes through the line segment between $ v_{1} $ and $
v_{n-1}, $ and these polygons form a codimension one subset of $
\mathfrak{Geo}^{(n)} .$ In particular, since $ \mathfrak{Geo}^{(n)} $ is a manifold, every
$ n $-sided polygon can be perturbed by just a tiny amount so that its
image under $ g $ lies in $ \mathfrak{Geo}^{(n-1)} . $
Suppose that $ Q $ is an $ (n-1) $-sided polygon in $ \mathfrak{Geo}^{(n-1)}. $
Then the preimage $ g^{-1}(Q) $ will be a three-dimensional manifold,
homeomorphic to the set of valid $ n $th vertices for $ Q. $ This
divides $ \mathfrak{Geo}^{(n)} $ into three-dimensional slices or {\em strata}.
As $ Q $ varies over $ \mathfrak{Geo}^{(n-1)}, $ the corresponding
three-dimensional stratus $ g^{-1}(Q) $ will vary. By observing how
these strata change, we can obtain useful information about $
\mathfrak{Geo}^{(n)} .$
\begin{figure}[t]
\insertfig{frontview.eps}{2.6in}
\caption{One possible sixth vertex for the pentagon $ Q .$}
\label{fig:frontview}
\end{figure}
For example, consider the pentagon $ Q = \langle v_1, v_2, v_3, v_4,
v_5 \rangle $ with coordinates
\begin{gather*}
\langle (0, 0, 0), (.886375, .276357, .371441), \\
(.125043, -.363873, .473812), \\
(.549367, .461959, .845227), (.818041, 0, 0) \rangle
\end{gather*}
shown in Figure ~\ref{fig:frontview}. Suppose that we replace the
edge between $ v_5 $ and $ v_1 $ with a pair of new edges, from $ v_5
$ to some new vertex $ v_{6} \in \Field{R}^3 $ and from this vertex back to $
v_1 .$ This creates a hexagon which, with a bit of care in choosing $
v_{6}, $ will also be embedded in $ \Field{R}^3 . $ For instance, if we
place the new vertex at $ (.4090205, 0, -.912525), $ we obtain an
unknotted hexagon. On the other hand, placing $ v_{6} $ at $
(.4090205, -.343939, .845227), $ gives a hexagon which is knotted as a
right-handed trefoil. See Figure ~\ref{fig:frontview}. The preimage
$ g^{-1}(Q) \in \mathfrak{Geo}^{(6)} $ is homeomorphic to the dense open subset
of $ \Field{R}^3 $ consisting of ``valid'' sixth vertices for $ Q .$
To examine which points in $ \Field{R}^3 $ correspond to embedded hexagons
obtained from $ Q, $ we will think of the $ x $-axis as a ``central
axis'' in this space and consider the collection of half-planes
radiating from this axis. We refer to these as {\em standard
half-planes}. These half-planes appear as rays from the origin in
Figure ~\ref{fig:sideview}, which shows the projection of $ Q $ into
the $ yz $-plane.
Let $ \mathcal{P}_{2}, \mathcal{P}_{3}, $ and $ \mathcal{P}_{4} $ be
the standard half-planes containing $ v_{2}, v_{3}, $ and $ v_{4}, $
respectively. Thus
\begin{gather*}
\mathcal{P}_{2} = \{ y = \tfrac{276357}{371441} z \approx .744 z,
\; z > 0 \} , \\
\mathcal{P}_{3} = \{ y = - \tfrac{363873}{473812} z \approx -.768 z,
\; z > 0 \} , \\
\mathcal{P}_{4} = \{ y = \tfrac{461959}{845227} z \approx .547 z,
\; z > 0 \} ,
\end{gather*}
as shown in Figure ~\ref{fig:sideview}.
\begin{figure}[t]
\insertfig{sideview.eps}{2.6in}
\caption{Projection of pentagon $ Q $ into $ yz $-plane.}
\label{fig:sideview}
\end{figure}
Notice that the interior of any standard half-plane to the left of $
\mathcal{P}_{2} $ and $ \mathcal{P}_{3} $ will miss $ Q $ altogether.
Thus, any point in the interior of any these half-planes may be used
as a sixth vertex for a hexagon. Every other standard half-plane,
however, does intersect $ Q $ at one or more interior points, so these
half-planes will contain some points which correspond to hexagons with
self-intersections.
The interior of any standard half-plane between $ \mathcal{P}_{2} $
and $ \mathcal{P}_{4} $ will intersect $ Q $ only once, in its second
edge. Depending on which point of this plane we choose for the new
vertex $ v_{6} , $ the two-edge linkage $ v_{5} v_{6} v_{1} $ will
either dip underneath or jump over this edge. If $ v_{5} v_{6} v_{1}
$ goes under $ v_{2} v_{3}, $ then $ v_{6} $ can be dragged back to
the $ x $-axis, say to the midpoint of edge $ v_{1}v_{5}, $ giving an
isotopy of the resulting hexagon back to the unknotted loop realized
by the pentagon $ Q .$ However, if $ v_{5} v_{6} v_{1} $ loops above
the edge $ v_{2}v_{3} , $ then this edge will obstruct any isotopy of
the hexagon which attempts to push $ v_{6} $ down towards the $ x
$-axis in this plane. For instance, $ Q $ crosses the half-plane $ \{
y = .6z, \; z > 0 \} $ at the point $ (.828333, .227547, .379246) . $
Vertices collinear with $ (.828333, .227547, .379246) $ and $ v_{1} $
correspond to embedded hexagons only when they lie between these
points; otherwise the second and sixth edges of the resulting hexagon
will cross each other. Similarly, vertices collinear with $ (.828333,
.227547, .379246) $ and $ v_{5} $ which do not lie between these two
points correspond to hexagons with intersecting second and fifth
edges. Therefore, points in the rays beginning at $ (.828333,
.227547, $ $ .379246) $ and radiating away from either $ v_{1} $ or $
v_{5} $ do not correspond to embedded hexagons, and the half-plane is
cut into two regions by a ``V''-shaped discriminant. See Figure
~\ref{fig:v.discrim}(a). If $ v_{6} $ is placed in the region of this
half-plane labelled {\em i}, then the pair of new edges will dip under
the edge $ v_{2}v_{3} $ and the resulting hexagon will be isotopic to
$ Q .$ Alternatively, if $ v_{6} $ is placed in the region labelled
{\em ii}, then $ v_{5} v_{6} v_{1} $ will jump over this edge.
\begin{figure}[t]
\insertfig{vdiscrim.eps}{1.7in}
\centering{{\small (a) $ \{ y = .6z, \; z > 0 \} $ \hspace{46mm}
(b) $ \{ y = 0, \; z > 0 \} $ \hspace{4mm}}}
\caption{$ Q $ separates each half-plane by ``V''-shaped discriminants.}
\label{fig:v.discrim}
\end{figure}
Now, the interior of every standard half-plane between $
\mathcal{P}_{4} $ and $ \mathcal{P}_{3} $ intersects $ Q $ in two
points, in the interior of its second and third edges. As before,
these edges will form obstructions to a homotopy moving $ v_{6} $ in
this plane. Therefore, for each of the points through which these
edges cross the half-plane, there will be a ``V''-shaped discriminant
as above. For example, in the half-plane $ \{ y = 0, \; z > 0 \} , $
which intersects $ Q $ at the points $ (.557744, 0, .415630) $ and $
(.312006, 0, .637463) , $ vertices in the four rays beginning at
either of these points and radiating away from $ v_{1} $ and $ v_{5} $
correspond to hexagons with self-intersections. These two
``V''-shaped discriminants separate the half-plane into four regions,
arranged as in Figure ~\ref{fig:v.discrim}(b). As before, placing the
new vertex in each of these regions corresponds to looping the new
edges of the hexagon over either the second ({\em vi}) or third ({\em
iv}) edge of $ Q , $ or both ({\em v}), or neither ({\em iii}) of
these.
We can show that the arrangement of the ``V''-shaped discriminants
remains relatively unchanged for standard half-planes in each of these
intervals. In fact, the connected components of the half-planes in
Figure ~\ref{fig:v.discrim} are only cross-sectional slices of
``cylindrical sectors'' of $ g^{-1}(Q) $ which wrap around the $ x
$-axis. Denote these sectors as {\em i}, {\em ii}, {\em iii}, {\em
iv}, {\em v}, and {\em vi}, using the notation in Figure
~\ref{fig:v.discrim}. Furthermore, let {\em o} denote the sector of $
g^{-1}(Q) $ corresponding to vertices in half-planes which do not
intersect $ Q $ at all. Then the way in which these sectors are glued
together depends on the behavior of the discriminants at the three
``critical level'' standard half-planes $ \mathcal{P}_{2} ,
\mathcal{P}_{3} , $ and $ \mathcal{P}_{4} .$
\begin{figure}[b]
\insertfig{p2discrim.eps}{1.54in}
\caption{Critical level $ \mathcal{P}_{2} = \{ y =
\tfrac{276357}{371441} z \approx .744 z, \; z > 0 \}.$ }
\label{fig:chp.discrim.a}
\end{figure}
The first of these half-planes, $ \mathcal{P}_{2}, $ contains the
first edge of $ Q , $ which connects $ v_{1} = (0,0,0) $ to $ v_{2} =
(.886375, .276357, .371441) . $ Vertices in rays beginning at any
point in this edge and radiating away from $ v_{6} $ correspond to
hexagons with intersecting first and fifth edges. Hence, for each
point in this edge there is a ``V''-shaped discriminant. The union of
these discriminants forms a two-dimensional discriminant corresponding
to an obstruction in the space $ g^{-1}(Q) .$ See Figure
~\ref{fig:chp.discrim.a}. However, this obstruction only
partially blocks access to sector {\em i}. Therefore both {\em i} and
{\em ii} are glued to {\em o} at this half-plane.
A similar two-dimensional discriminant occurs for $ \mathcal{P}_{4} .$
This half-plane contains the fourth edge of $ Q , $ which joins $
v_{4} = (.549367, .461959, .845227) $ and $ v_{5} = (.818041, $ $ 0,
0).$ Vertices collinear with the origin and any point $ p $ on this
edge correspond to embedded hexagons only if they lie between $ (0, 0,
0) $ and $ p .$ See Figure ~\ref{fig:chp.discrim.b}. This
discriminant completely closes off sector {\em vi}, and obstructs
parts of sectors {\em i}, {\em ii}, and {\em iii}. Thus, at this
level, {\em i} is attached to {\em iii} and {\em iv}, {\em ii} is
attached to {\em v}, and {\em vi} is abruptly terminated.
\begin{figure}[h]
\insertfig{p4discrim.eps}{2.13in}
\caption{Critical level $ \mathcal{P}_{4} = \{ y =
\tfrac{461959}{845227} z \approx .547 z, \; z > 0 \}.$}
\label{fig:chp.discrim.b}
\end{figure}
The third critical level half-plane, $ \mathcal{P}_{3}, $ presents a
different situation, as it intersects $ Q $ only at the vertex $ v_{3}
= (.125043, -.363873, .473812) .$ In this case, the ``V''-shaped
discriminants corresponding to the second and third edges of $ Q $
come together as the two edges become incident at their common vertex.
As the discriminants merge, sectors {\em iv} and {\em vi} are
terminated, while both of the sectors {\em iii} and {\em v} merge with
sector {\em o}.
Figure ~\ref{fig:cylinder} presents a cylindrical section of $ \Field{R}^3 $
about the $ x $-axis, showing the sectors of $ g^{-1}(Q) $ and the
connections between them. In particular, it shows that $ g^{-1}(Q) $
consists of two disjoint path-components, corresponding to the two
knot types possible for hexagons in the stratus $ g^{-1}(Q) $: the
unknot and the right-handed trefoil.
\begin{figure}[t]
\insertfig{cylinder.eps}{2in}
\caption{Cylindrical section of $ \Field{R}^{3} $ showing the different
interconnecting sectors of $ g^{-1}(Q) .$}
\label{fig:cylinder}
\end{figure}
The key feature characterizing a pentagon's corresponding stratus in $
\mathfrak{Geo}^{(6)} $ is the relative position of the second, third, and fourth
vertices with respect to the axis through the other two vertices.
Suppose that $ Q = \langle v_1, v_2, v_3, v_4, v_5 \rangle $ is an
arbitrary pentagon in $ \mathfrak{Geo}^{(5)}, $ and that $ \mathcal{L} $ is the
line determined by $ v_1 $ and $ v_5 .$ Since $ \mathfrak{Geo}^{(5)} $ is a
manifold, we can perturb $ Q $ slightly, if necessary, to ensure that
$ v_1 v_5 $ is the only edge of $ Q $ which intersects $ \mathcal{L}.$
As above, let $ \mathcal{P}_{2}, \mathcal{P}_{3}, $ and $
\mathcal{P}_{4} $ be the half-planes with boundary $ \mathcal{L} $
which contain $ v_2, v_3, $ and $ v_4 , $ respectively. Again, a
slight deformation will make $ Q $ a generic pentagon, guaranteeing
that the three $ \mathcal{P}_{i} $'s are distinct.
As in the example above, the $ \mathcal{P}_{i} $'s will divide $
\Field{R}^{3} $ into three open regions, with $ Q $ intersecting two of these
and completely missing the third. As we rotate in a right-handed
fashion about the axis $ \mathcal{L} , $ beginning in the region which
misses $ Q , $ we will encounter each of the $ \mathcal{P}_{i} $'s in
one of six orders. For example, in the pentagon shown in Figures
~\ref{fig:frontview} and \ref{fig:sideview}, these half-planes appear
in the order $ \mathcal{P}_2 - \mathcal{P}_4 - \mathcal{P}_3 , $ or
simply, 2-4-3.
Let $ H \in \mathfrak{Geo}^{(6)} $ be a generic hexagon embedded in $ \Field{R}^{3} .$
By considering the order in which the $ \mathcal{P}_{i} $'s associated
with $ g(H) $ occur, we divide $ \mathfrak{Geo}^{(6)} $ into six open regions
meeting along codimension one sets where either:
\begin{enumerate} \renewcommand{\labelenumi}{(\roman{enumi})}
\item two of the $ \mathcal{P}_{i} $'s coincide, or
\item an edge of $ H $ crosses $ \mathcal{L} .$
\end{enumerate}
By analyzing the behavior of the strata $ g^{-1}(g(H)) $ as they
interconnect, we obtain Table ~\ref{tab:hexagons}, which indicates the
number of path-components in each of the six regions of $ \mathfrak{Geo}^{(6)} ,
$ arranged by the topological knot type they represent.
\begin{table}[ht]
\caption{Number of components in each region of $ \mathfrak{Geo}^{(6)} .$ }
\label{tab:hexagons}
\begin{center}
\begin{tabular}{|ccccc|}
\hline
$ \mathbf{Region \, of \, \mathfrak{Geo}^{(6)}} $ & \; &
$ \quad \mathbf{0} \quad $ &
$ \quad \mathbf{3_1} \quad $ &
$ \; \, \mathbf{- 3_1} \; \, $ \\
\hline
2-3-4 && 1 & - & - \\
2-4-3 && 1 & 1 & - \\
3-2-4 && 1 & 1 & - \\
3-4-2 && 1 & - & 1 \\
4-2-3 && 1 & - & 1 \\
4-3-2 && 1 & - & - \\
\hline
\end{tabular}
\end{center}
\end{table}
As noted above, the six regions of $ \mathfrak{Geo}^{(6)} $ meet along
codimension one subsets consisting of hexagons for which two of the $
\mathcal{P}_i $'s coincide. For instance, regions 2-4-3 and 4-2-3
meet along a subset consisting of hexagons with $ \mathcal{P}_2 =
\mathcal{P}_4 .$ The six regions also meet along codimension one
subsets consisting of hexagons that intersect line $ \mathcal{L} . $
For example, regions 2-4-3 and 4-3-2 meet along a set of hexagons for
which edge $ v_2 v_3 $ intersects this line. These connections are
shown schematically in Figure~\ref{fig:connections}; here solid lines
represent hexagons with two coinciding $ \mathcal{P}_i $'s while gray
lines represent hexagons for which some edge intersects line $ v_1 v_5
.$
\begin{figure}[t]
\insertfig{connect.eps}{2in}
\caption{Codimension one connections between regions.}
\label{fig:connections}
\end{figure}
Consider a hexagon $ H $ in the common boundary between two regions of
$ \mathfrak{Geo}^{(6)} .$ Since $ H $ can be perturbed slightly to make generic
hexagons of either type, $ H $ must be of a topological knot type
common to both regions. However, the only knot type common to
adjacent regions in Figure~\ref{fig:connections} is the unknot.
Therefore hexagons in these codimension one subsets must be unknotted
and, in particular, the topological unknots form a single component of
geometric unknots in $ \mathfrak{Geo}^{(6)} .$
On the other hand, suppose that $ h: [0, 1] \rightarrow \mathfrak{Geo}^{(6)} $
is a path from some trefoil of type 2-4-3 to some trefoil of type
3-2-4. Since $ \mathfrak{Geo}^{(6)} $ is an open subset of $ \Field{R}^{18} , $ there
is a small open 18-ball contained in $ \mathfrak{Geo}^{(6)} $ about each point
in this path. Thus we can assume that whenever $ h $ passes through a
boundary of one of the six regions, it does so through a generic point
in one of the codimension one subsets above. But then $ h $ must pass
through either 2-3-4 or 4-3-2; see Figure~\ref{fig:connections}. This
is a contradiction since only unknots live in these regions. Thus
there is no path connecting the trefoils of type 2-4-3 and those of
type 3-2-4. Similarly, there is no path between the type 4-2-3 and
3-4-2 trefoils. This proves that $ \mathfrak{Geo}^{(6)} $ consists of five
path-components: one consisting of unknots, two of right-handed
trefoils, and two of left-handed trefoils.
The geometric knot types in $ \mathfrak{Geo}^{(6)} $ are completely
characterized by a pair of combinatorial invariants which capture a
hexagon's topological chirality ({\em i.e.} right- or left-handedness)
and geometric curl ({\em i.e.} ``upward'' or ``downward'' twisting),
and are easily computed from the coordinates of a hexagon's vertices.
To define these invariants, let $ H = \langle v_{1}, v_{2}, v_{3},
v_{4}, v_{5}, v_{6} \rangle $ be an embedded hexagon in $ \Field{R}^{3} ,$
and consider the open triangular disc determined by vertices $ v_{1},
v_{2}, $ and $ v_{3}.$ This disc inherits an orientation from $ H $
via the ``right hand rule.'' Let $ \Delta_{2} $ be the algebraic
intersection number of the hexagon and this triangle. Notice that the
triangular disc can only be pierced by edges $ v_{4}v_{5} $ and $
v_{5}v_{6}.$ Furthermore, if both of these edges intersect the disc,
they will do so in opposite directions, with their contributions to $
\Delta_{2} $ canceling out. Thus $ \Delta_{2} $ takes on a value of
0, 1, or $ -1 .$ Similarly, define $ \Delta_{4} $ and $ \Delta_{6} $
to be the intersection numbers of H with the triangles $ \triangle
v_{3}v_{4}v_{5} $ and $ \triangle v_{5}v_{6}v_{1}, $ respectively. By
considering the possible values for the $ \Delta_{i} $'s (see Lemma 8
in \cite{Calvo:hexagons}), we can show that
\begin {enumerate} \renewcommand {\labelenumi} {(\roman{enumi})}
\item $ H $ is a right-handed trefoil if and only if $ \Delta_2 =
\Delta_4 = \Delta_6 = 1 , $
\item $ H $ is a left-handed trefoil if and only if $ \Delta_2 =
\Delta_4 = \Delta_6 = -1 , $ and
\item $ H $ is an unknot if and only if $ \Delta_i = 0 $ for some $ i
\in \{2, 4, 6 \} ,$
\end{enumerate}
implying that the product
\begin{equation} \label{eq:chirality}
\Delta(H) = \Delta_{2}\Delta_{4}\Delta_{6},
\end{equation}
which we call the {\em chirality} of $ H $, is an invariant under
geometric deformations.
Next, we define the {\em curl} of $ H $ as
\begin{equation} \label{eq:curl}
\Curl H = \Sign \bigl( (v_3 - v_1) \times (v_5 - v_1) \cdot (v_2 -
v_1) \bigr) .
\end{equation}
This gives the sign of the $ z $-coordinate of $ v_2 $ when we rotate
$ H $ so that $ v_1, v_3, $ and $ v_5 $ are placed on the $ xy $-plane
in a counterclockwise fashion, and therefore measures in some sense
whether a hexagon twists up or down. Consider a path $ h: [0, 1]
\rightarrow \mathfrak{Geo}^{(6)} $ which changes the curl of a hexagonal trefoil
from +1 to -1. Then there must be some point on this path for which
the vector triple product in (\ref{eq:curl}) is equal to zero. At
this point, the vertices $ v_1, v_2, v_3, $ and $ v_5 $ are all
coplanar. However, we can show such a hexagon must be unknotted,
giving us a contradiction. In particular, the product $ \Delta^{2}(H)
\, \Curl H $ is also an invariant under geometric deformations. A
simple calculation then shows that every trefoil of type 2-4-3 or
4-2-3 has positive curl, while every one of type 3-2-4 or 3-4-2 has
negative curl.
\begin{Thm} \label{thm:jcc}
Define the joint chirality-curl of a hexagon $ H $ as the ordered pair
$ \mathcal{J} (H) = (\Delta(H), \Delta^{2}(H) \, \Curl H) .$ Then
\begin{equation} \label{eq:jcc.values}
\mathcal{J} (H) = \begin{cases} ( 0, 0 ) & \text{iff} \, H \, \text{is an
unknot,} \\
( +1, c ) & \text{iff} \, H \, \text{is a right-handed trefoil with}
\, \Curl H = c, \\
( -1, c ) & \text{iff} \, H \, \text{is a left-handed trefoil with} \,
\Curl H = c.
\end{cases}
\end{equation}
Therefore the geometric knot type of a hexagon $ H $ is completely
determined by the value of its chirality and curl.
\end{Thm}
Before leaving the world of hexagons behind, let us make one last
observation. Recall that the construction of $ \mathfrak{Geo}^{(n)} $ depends
on a choice of a ``first vertex'' $ v_1 $ and an orientation. This
amounts to choosing a sequential labeling $ v_1, v_2, \ldots, v_n $
for the vertices of each polygon. A different choice of labels will
lead to a different point in $ \mathfrak{Geo}^{(n)} $ corresponding to the same
underlying polygon. Thus the dihedral group $ \mathbf{D}_n $ of order
$ 2n $ acts on $ \mathfrak{Geo}^{(n)} $ by shifting or reversing the order of
these labels, and this action preserves topological knot type.
Observation of the effects on $ \Curl H $ by the group action of $
\mathbf{D}_6 $ on $ \mathfrak{Geo}^{(6)} $ reveals that same statement does not
hold true for geometric knot type. In particular, if the group action
is defined by the automorphisms
\begin{align*}
r : \langle v_1, v_2, v_3, v_4, v_5, v_6 \rangle & \mapsto \langle v_1, v_6,
v_5, v_4, v_3, v_2 \rangle , \\
s : \langle v_1, v_2, v_3, v_4, v_5, v_6 \rangle & \mapsto \langle v_2, v_3,
v_4, v_5, v_6, v_1 \rangle ,
\end{align*}
then
\[ \Curl r H = \Curl s H = - \Curl H . \]
This shows that the hexagonal trefoil knot is not reversible: In
contrast with trefoils in the topological setting, reversing the
orientation on a hexagonal trefoil yields a different geometric knot.
Furthermore, shifting the labels over by one vertex also changes the
the knot type of a trefoil, so that taking quotients under this action
we can see that the spaces $ \mathfrak{Geo}^{(6)} / \! \prec \! s \! \succ $
of non-based, oriented hexagons, and $ \mathfrak{Geo}^{(6)} / \! \prec \! r, s
\! \succ $ of non-based, non-oriented hexagons consist of only three
components each.
A similar decomposition can be made for the space $ \mathfrak{Geo}^{(7)} $ of
heptagons. In this case we consider the relative ordering of the
half-planes $ \mathcal{P}_{2}, \mathcal{P}_{3}, \mathcal{P}_{4}, $ and
$ \mathcal{P}_{5} $ bounded by the line through $ v_{1} $ and $ v_{6}.
$ This defines 24 open regions which meet along codimension one
subsets where two of the $ \mathcal{P}_i $'s coincide. These
junctions can be schematically described as switches in the indices
denoting the regions. For instance, regions 2-4-3-5 and 4-2-3-5 meet
along a subset consisting of hexagons with $ \mathcal{P}_2 =
\mathcal{P}_4 . $ We can build a model for these connections by
taking a vertex for each of the 24 regions and an edge for each
codimension-1 subset joining them. The result is a valence-3 graph
which forms the 1-dimensional skeleton of a solid zonotope called a
{\em permutahedron}, shown in Figure ~\ref{fig:permutahedron}. Each
vertex of the permutahedron is part of a unique square face
corresponding to the order-4 sequence of index switches in which the
first two indices and the last two indices are switched in an
alternating fashion. In addition, each vertex is part of two distinct
hexagonal faces which correspond to the order-6 switch sequences in
which either the first or last index is fixed while the other three
indices are permuted through all six possible orderings. Therefore
the valence-3 permutahedron has six square faces and eight hexagonal
faces. Extending the edges shared by any two hexagonal faces shows
that this is nothing more than a truncated octahedron, also known in
crystallography as a Fedorov cubo-octahedron. ~\!\!\footnote{\, The
cubo-octahedron is a {\em parallelohedron}, that is, a crystalline
shape having parallel opposite faces with which three-space can be tiled.
One should not confuse Fedorov's cubo-octahedron with Kepler's
cuboctahedron, which is built from an octahedron by truncating at the
midpoint (rather than at the one- and two-third points) of each edge
and thus consists of six squares and eight triangular faces. See
pp.17 -- 18 in \cite{Ziegler:polytopes} and pp. 722 -- 723 in
\cite{Tutton:crystals}.}
\begin{figure}[t]
\insertfig{permuta.eps}{2in}
\caption{The valence 3 permutahedron.}
\label{fig:permutahedron}
\end{figure}
With a few additional considerations, ~\!\!\footnote{\, The interested
reader is referred to pp. 53 -- 56 in \cite{Calvo:thesis}.} the
analysis of the strata over each of these regions shows that $
\mathfrak{Geo}^{(7)} $ has a single path-component of unknots and of each
topological type of trefoil, and two containing figure-eight knots.
Again, these figure-eight knots are new examples of distinct geometric
isotopes of the same topological knot, which can be distinguished by a
geometric invariant $ \Xi, $ defined as follows.
Suppose that $ H $ is the heptagon $ \langle v_1, v_2, v_3, v_4, v_5,
v_6, v_7 \rangle. $ Define the functions $ \Theta_3 (H) $ and $
\Theta_6 (H) $ as
\begin{equation}
\begin{aligned}
\Theta_3 &= \Sign \bigl( (v_7 - v_1) \times (v_2 - v_1) \cdot (v_3 - v_1) \bigr), \\
\Theta_6 &= \Sign \bigl( (v_6 - v_1) \times (v_7 - v_1) \cdot (v_2 - v_1) \bigr).
\end{aligned}
\end{equation}
Then $ \Theta_3 = \Theta_6 $ if the vertices $ v_3 $ and $ v_6 $ lie
on the same side of the plane $ \mathcal{P} $ determined by $ v_7,
v_1, $ and $ v_2 , $ and $ \Theta_3 = - \Theta_6 $ if $ v_3 $ and $
v_6 $ lie on different sides of $ \mathcal{P} . $ Notice that
for a generic heptagon, exactly one of the functions $ \frac{1}{2}
(\Theta_3 + \Theta_6) $ and $ \frac{1}{2} (\Theta_3 - \Theta_6) $ is
zero, while the other is $ \pm 1 . $
Let $ I_{34} $ denote the algebraic intersection number of edge $ v_3
v_4 $ with the triangular disc $ \triangle v_{7}v_{1}v_{2}, $ using
the usual orientations induced by $ H . $ Similarly define $ I_{45} $
and $ I_{56} $ as the intersection numbers of the triangle $ \triangle
v_{7}v_{1}v_{2} $ with the edges $ v_4 v_5 $ and $ v_5 v_6 , $
respectively.
If $ H $ has $ \Theta_3 = \Theta_6 , $ then $ v_3 $ and $ v_6 $ lie on
the same side of the plane $ \mathcal{P} $ so that the three-edge
linkage $ v_3 v_4 v_5 v_6 $ will intersect $ \mathcal{P} $ at most
twice. Furthermore, if both of these intersections happen in the
interior of $ \triangle v_{7}v_{1}v_{2}, $ they occur with opposite
orientations. Thus the sum $ I_{34} + I_{45} + I_{56} $ only takes on
values -1, 0, or 1.
On the other hand, suppose that $ H $ is a figure-eight knot with $
\Theta_3 = - \Theta_6 . $ Then $ v_3 $ and $ v_6 $ lie on the
opposite sides of the plane $ \mathcal{P} $ so that the three-edge
linkage $ v_3 v_4 v_5 v_6 $ intersects $ \mathcal{P} $ an odd number
of times. First, suppose that there is only one intersection; then
the linkage $ v_7 v_1 v_2 $ can be piecewise linearly isotoped into a
straight line segment. We can think of this isotopy as either pushing
$ v_{1} $ in a straight line path towards the midpoint of the line
segment $ v_{2} v_{7}, $ or (in the case that the intersection occurs
inside $ \triangle v_{7}v_{1}v_{2} $) as stretching $ v_7 v_1 v_2 $
into a large loop, swinging it like a ``jump rope'' around and to the
other side of the heptagon, and then pushing it in until it coincides
with the line segment $ v_{2} v_{7} . $ See Figure
~\ref{fig:jumprope}. In either case we get a hexagonal realization of
a figure-eight knot. Since this is impossible, the linkage $ v_3 v_4
v_5 v_6 $ has to cross the plane $ \mathcal{P} $ three times, and in
particular, $ v_3 v_4 $ and $ v_5 v_6 $ must do so with the same
orientation. Therefore the quantity $ I_{34} - I_{56} $ will either
be zero (when both edges intersect $ \triangle v_{7}v_{1}v_{2}, $ or
when neither of the two do) or $ \pm 1 $ (when only one of these
intersections occurs inside the triangle).
\begin{figure}[t]
\insertfig{jumprope.eps}{2.75in}
\caption{A piecewise linear isotopy of the linkage $ v_7 v_1 v_2 .$}
\label{fig:jumprope}
\end{figure}
A quick look at the possible configurations shows that:
\begin{enumerate} \renewcommand {\labelenumi} {(\roman{enumi})}
\item if $ H $ is a heptagonal figure-eight knot with $ \Theta_3 =
\Theta_6 , $ then exactly one of the intersection numbers $ I_{34},
I_{45}, $ or $ I_{56} $ is non-zero; thus $ I_{34} + I_{45} + I_{56} =
\pm 1 $ (Lemma 4.2 in \cite{Calvo:thesis}), and
\item if $ H $ is a heptagonal figure-eight knot with $ \Theta_3 = -
\Theta_6 , $ then exactly one of the intersection numbers $ I_{34} $
or $ I_{56} $ is non-zero; in particular $ I_{34} - I_{56} = \pm 1 $
(Lemma 4.3 in \cite{Calvo:thesis}).
\end{enumerate}
Now consider the function
\begin{equation} \label{eq:xi}
\Xi (H) = \frac{1}{2} \biggl( \Theta_3 + \Theta_6 \biggr)
\biggr( I_{34} + I_{45} + I_{56} \biggr) +
\frac{1}{2} \biggl( \Theta_3 - \Theta_6 \biggr)
\biggr( I_{34} - I_{56} \biggr) .
\end{equation}
By (i) and (ii) above, $ \Xi $ can only take values of 1 or -1.
Suppose that the value of $ \Xi $ changes along some path $ h: [0,1]
\rightarrow \mathfrak{Geo}^{(7)} .$ Since $ \mathfrak{Geo}^{(7)} $ is a manifold, we can
assume that, in this path, only one vertex passes through the interior
of $ \triangle v_{7}v_{1}v_{2} $ at any one time, and that only one
edge intersects the line segment $ v_7 v_2 $ at a time, and that these
two things happen at different times. Note that each of these events
will change the values of $ I_{34} + I_{45} + I_{56} $ and $ I_{34} -
I_{56} $ by at most one. However, $ \Xi $ can only change in
increments of two, so if the values of $ \Theta_3 $ and $ \Theta_6 $
remain constant through out $ h, $ $ \Xi $ must also remain unchanged.
By reversing orientations if necessary, we can assume then that the
deformation changes the sign of $ \Theta_3 . $ In particular, let $
H_0 $ be a heptagonal figure-eight knot with $ \Theta_3 = 0 . $ By
pushing $ v_3 $ slightly towards $ v_6 , $ we get a heptagon $ H_0^+ $
with $ \Theta_3 = \Theta_6 ; $ let $ I_{34}^+ $ be the appropriate
intersection number for this heptagon. On the other hand, we obtain a
heptagon $ H_0^- $ with $ \Theta_3 = - \Theta_6 $ by pushing $ v_3 $
away from $ v_6 ; $ let $ I_{34}^- $ be the corresponding intersection
number for this heptagon. By picking $ H_{0}^{+} $ and $ H_{0}^{-} $
close enough to $ H_{0}, $ we can assume that the values of the
intersection numbers $ I_{45} $ and $ I_{56} $ coincide for all three
knots. This leaves two cases to consider.
First, suppose that $ I_{34}^- = 0 . $ Then $ I_{56} = \pm 1 $ by
(ii), and hence $ I_{34}^+ = I_{45} = 0 $ by (i). Therefore
\[ \biggl( I_{34}^+ + I_{45} + I_{56} \biggr) = I_{56} =
- \biggl( I_{34}^- - I_{56} \biggr) . \]
The extra negative sign in the right hand term of this equation
neutralizes the change of sign in $ \Theta_{3}, $ so that $ \Xi $
remains unchanged.
Next, suppose that $ I_{34}^- = \pm 1, $ in which case $ I_{34}^+ = 0
.$ Then $ I_{56} = 0 $ by (ii) and $ I_{45} = \pm 1 $ by (i).
Furthermore, the edges $ v_3 v_4 $ and $ v_4 v_5 $ must intersect the
interior of $ \triangle v_{7}v_{1}v_{2} $ from opposite directions, so
$ I_{34}^- = - I_{45} .$ Therefore
\[ \biggl( I_{34}^+ + I_{45} + I_{56} \biggr) = I_{45} = - I_{34}^- =
- \biggl( I_{34}^- - I_{56} \biggr) , \]
so that, as before, $ \Xi $ does not change. This proves the
following result.
\begin{Thm}
$ \Xi $ is an invariant of heptagonal figure-eight knots under geometric
deformations.
\end{Thm}
It is interesting to note that $ \Xi $ is also invariant under mirror
reflections, since the resulting sign changes in the functions $
\Theta_{3}, \Theta_{6}, I_{34}, I_{45}, $ and $ I_{56} $ cancel out in
(\ref{eq:xi}). This reflects the fact that heptagonal figure-eight
knots are achiral, {\em i.e.} equivalent to their mirror images.
Figure ~\ref{fig:achiral} shows one such isotopy. Starting with the
diagram at the top of Figure ~\ref{fig:achiral} and proceeding in a
clockwise fashion, we first push $ v_1 $ through the interior of the
triangular disc $ \triangle v_2 v_3 v_4 .$ Note that in doing so, we
may need to change the lengths of one or more of the edges. Although
it is difficult to see from the perspective of Figure
~\ref{fig:achiral}, this motion actually defines an isotopy from the
heptagon $ \langle v_1, v_2, v_3, v_4, v_5, v_6, v_7 \rangle $ to the
heptagon $ \langle -v_6, -v_7, -v_1, -v_2, -v_3, -v_4, -v_5 \rangle .$
We continue by repeating similar moves, passing $ v_3 $ through $
\triangle v_4 v_5 v_6 , $ then $ v_5 $ through $ \triangle v_6 v_7 v_1
, $ and so on. After seven steps, when we move $ v_6 $ past $
\triangle v_7 v_1 v_2 , $ we arrive at the diagram at the bottom of
Figure ~\ref{fig:achiral}. At this point, the figure-eight knot is
the mirror image of the starting position.
\begin{figure}[p]
\vspace{.5in}
\insertfig{achiral.eps}{6.4in}
\caption{Heptagonal figure-eight knots are achiral.}
\label{fig:achiral}
\end{figure}
Finally, consider the $ \mathbf{D}_{7} $ action on $ \mathfrak{Geo}^{(7)} $ defined by
the automorphisms
\begin{align*}
r \langle v_1, v_2, v_3, v_4, v_5, v_6, v_7 \rangle &=
\langle v_1, v_7, v_6, v_5, v_4, v_3, v_2 \rangle \\
s \langle v_1, v_2, v_3, v_4, v_5, v_6, v_7 \rangle &=
\langle v_2, v_3, v_4, v_5, v_6, v_7, v_1 \rangle .
\end{align*}
Reversing the orientation on $ H $ via the map $ r $ will reverse the
orientations on both the edges of $ H $ and the triangular discs that
they define. In particular,
\[ I_{34}(rH) = I_{56}(H) \qquad \qquad I_{45}(rH) = I_{45}(H)
\qquad \qquad I_{56}(rH) = I_{34}(H) .
\]
On the other hand, $ r $ not only switches the roles of $ \Theta_3 $ and $ \Theta_6, $
but also changes their signs:
\begin{align*}
\Theta_3 (rH) &= \Sign \bigl( (v_2 - v_1) \times (v_7 - v_1) \cdot (v_6 - v_1) \bigr), \\
&= - \Sign \bigl( (v_6 - v_1) \times (v_7 - v_1) \cdot (v_2 - v_1) \bigr), \\
&= - \Theta_6 (H), \\
\Theta_6 (rH) &= \Sign \bigl( (v_3 - v_1) \times (v_2 - v_1) \cdot (v_7 - v_1) \bigr), \\
&= - \Sign \bigl( (v_7 - v_1) \times (v_2 - v_1) \cdot (v_3 - v_1) \bigr), \\
&= - \Theta_3 (H). \\
\end{align*}
Therefore
\begin{align*}
\Xi (rH) &= \frac{1}{2} \biggl( \Theta_3 (rH) + \Theta_6 (rH) \biggr)
\biggr( I_{34} (rH) + I_{45} (rH) + I_{56} (rH) \biggr) \\ & \hspace{25mm} +
\frac{1}{2} \biggl( \Theta_3 (rH) - \Theta_6 (rH) \biggr)
\biggr( I_{34} (rH) - I_{56} (rH) \biggr) \\
&= \frac{1}{2} \biggl( - \Theta_6 (H) - \Theta_3 (H) \biggr)
\biggr( I_{56} (H) + I_{45} (H) + I_{34} (H) \biggr) \\ & \hspace{30mm} +
\frac{1}{2} \biggl( - \Theta_6 (H) + \Theta_3 (H) \biggr)
\biggr( I_{56} (H) - I_{34} (H) \biggr) \\
&= - \Xi (H) .
\end{align*}
This shows that, like hexagonal trefoil knots, figure-eight knots in $
\mathfrak{Geo}^{(7)} $ are irreversible, in contrast with their topological
counterparts. However, recall that the irreversibility of trefoils in
$ \mathfrak{Geo}^{(6)} $ depended strongly on our choice of a ``first'' vertex $
v_1 .$ In that case, a cyclic permutation of its six vertices would
change the trefoil's geometric knot type. This is not the case for
the figure-eight knots in $ \mathfrak{Geo}^{(7)}, $ for consider the group
action induced by the automorphism $ s $ on the set of geometric
isotopes of the figure-eight knot. This is an order 7 action on a two
element set, and must therefore be trivial. In other words, we must
have
\[ \Xi (sH) = \Xi (H). \]
Hence the distinction in the two figure-eight knot types is an effect
of ``true'' geometric knotting, which goes beyond a simple relabeling
of the vertices or our arbitrary choice of first vertex.
\section{Knot Projections and Minimal Polygon index} \label{sec:project}
In Section ~\ref{sec:stratify}, we were concerned with the question of
determining, for a given integer $ n, $ the number of path-components
present in the space $ \mathfrak{Geo}^{(n)} $ of $ n $-sided polygons. In other
words, ``how many geometric knot types are there for a particular
value of $ n $?'' However, as $ n $ increases, the space $ \mathfrak{Geo}^{(n)}
$ becomes more and more combinatorially intricate. As this happens,
we turn to the question of understanding the number of represented
topological (rather than geometric) knot types, and in particular, of
how complicated a knot can be realized by an $ n $-sided polygon. The
answer to this question is only known when $ n \le 8 .$ For example,
we know there are 9-sided polygonal embeddings of every seven crossing
prime knot $ ( 7_{1}, \ldots , 7_{7}) $ as well as the knots $ 8_{16},
8_{17}, 8_{18}, 8_{21}, 9_{40}, 9_{41}, 9_{42}, $ and $ 9_{46}, $ but
presumably this list could be much bigger, and include some of the
knots for which we have so far only found 10- or 11-sided
realizations.~\!\!\footnote{\, See Table 1 in
\cite{Calvo:montecarlo}.} In this section, we give one of several
known bounds on the complexity of an $ n $-sided polygon.
Recall that the {\em minimal crossing number} of a knot is the
smallest number of crossings present in any general position
projection of the knot into a plane or sphere. This is the
conventional measure of a knot's complexity, used in the standard
notation for knots and links as well as in the knot tables in the
appendices of \cite{Adams:book}, \cite{Kauffman:book},
\cite{Livingston:book}, and \cite{Rolfsen:book}. We similarly define
the {\em minimal polygon index} of a knot as the smallest number of
edges present in any polygonal embedding of the knot. This invariant,
which is elsewhere known as the {\em stick number} \cite{Adams:book,
Adams:sticks, Furstenberg:sticks}, the {\em broken line number}
\cite{Negami:s.bounds}, or simply the {\em edge number}
\cite{Meissen:sticks, Randell:sticks}, serves as the corresponding
measure of complexity for polygonal knots. These two invariants are
traditionally related by the following construction.
~\!\!\footnote{\, This construction appears in Theorem 7 in
\cite{Negami:s.bounds}, and as Exercise 1.38 in \cite{Adams:book}.}
Let $ P = \langle v_1, v_2, \ldots, v_{n-1}, v_n \rangle \in
\mathfrak{Geo}^{(n)} $ be an $ n $-sided polygon embedded in $ \Field{R}^{3} .$ We
project the points in $ P $ orthogonally onto a plane perpendicular to
one of its edges, say $ v_{1} v_{2} .$ This amounts to looking at the
polygon from a viewpoint in which we see the edge $ v_{1}v_{2} $
``head on,'' so that the image of $ P $ on our retina ({\em i.e.} the
plane) is an $ (n-1) $-sided polygon. An edge in this image cannot
cross either of its two neighbors, or itself, so each edge will
intersect at most $ n-4 $ other edges. Thus for generic polygons in $
\mathfrak{Geo}^{(n)}, $ this method gives a knot projection with no more than $
\tfrac{1}{2} (n-1)(n-4) $ crossings. This leads to the conclusion
that if a knot $ K $ has minimal crossing number $ c(K) $ and minimal
polygon index $ s(K) ,$ then \[ c(K) \le \frac{\bigl( s(K) - 1 \bigr)
\bigl( s(K) - 4 \bigr)}{2} , \] or equivalently (by completing squares
and solving for $ s $) \[ s(K) \ge \frac{5 \, + \, \sqrt{9 \, + \, 8
c(K)}}{2} .\]
For hexagons and heptagons, the bound on crossing number becomes 5 and
9, respectively, well over the actual values of 3 and 4 obtained in
Section ~\ref{sec:stratify}. In fact, the estimated $ \tfrac{1}{2}
(n-1)(n-4) $ crossings in the image of an $ n $-sided polygon can
never be achieved when $ n $ is odd. Here we present an improvement
on the bounds above.
First suppose that we relabel the vertices of $ P $ in sequence so
that $ v_{1} $ is a point on the boundary of the convex hull spanned
by the vertices of $ P .$ Therefore, we can find a plane $
\mathcal{P}_{1} $ which intersects $ P $ only at the vertex $ v_{1}, $
with $ P $ lying entirely on one side of $ \mathcal{P}_{1} .$
Let $ \mathcal{S} $ be a large sphere centered at $ v_1 $ and
enclosing all of $ P , $ and consider the image of the radial
projection $ p: P - \{v_1\} \rightarrow \mathcal{S} .$ By our choice
of $ v_{1} ,$ this image lies entirely in a hemisphere of $
\mathcal{S} $ cut by the equator $ \mathcal{S} \cap \mathcal{P}_{1} .$
Furthermore, note that the interiors of edges $ v_1 v_2 $ and $ v_1
v_n $ are respectively mapped to the single points $ p(v_2) $ and $
p(v_n) .$ Thus, by picking a generic $ P $ in $ \mathfrak{Geo}^{(n)} , $ we can
assume that $ \Gamma = p (P - \{v_1\}) $ consists of a chain of $ n-2
$ great circular arcs on $ \mathcal{S} $ intersecting in four-valent
crossings.
Suppose that $ \Gamma $ has $ c $ crossings. Since $ \Gamma $ is
contained in a single hemisphere of $ \mathcal{S}, $ a pair of arcs
will intersect at most once. Furthermore adjacent arcs cannot
intersect, so each one of the $ n-4 $ interior arcs $ p(v_3 v_4),
\ldots, p(v_{n-2} v_{n-1}) $ can intersect at most $ n-5 $ other arcs,
while each of the extreme arcs $ p(v_2 v_3) $ and $ p(v_{n-1} v_n) $
can intersect at most $ n-4 $ other arcs. Hence
\[
c \le \frac{1}{2}\biggl( (n-4)(n-5) + 2(n-4) \biggr) = \frac{(n-3)(n-4)}{2}.
\]
\begin{figure}[t]
\insertfig{epsilon1.eps}{2in}
\centering{{\small (a) \hspace{48mm} (b) \hspace{4mm}}}
\insertfig{epsilon2.eps}{2in}
\centering{{\small (c) \hspace{48mm} (d) \hspace{4mm}}}
\caption{Deformation of $ P $ inside a small $ \epsilon $-ball
about $ v_{1}.$}
\label{fig:epsilon.ball}
\end{figure}
Let $ \epsilon > 0 $ be small enough that the closed $ \epsilon $-ball
$ \mathcal{B}_{\epsilon} $ centered at $ v_{1} $ intersects the
polygon $ P $ in exactly two small segments of the edges $ v_{1}v_{2}
$ and $ v_{n}v_{1} , $ as shown in Figure ~\ref{fig:epsilon.ball}(a).
Suppose that the edge $ v_{1}v_{2} $ intersects the sphere $ \partial
\mathcal{B}_{\epsilon} $ at the point $ q_{1} .$ Furthermore, let $
q_{2} $ be the point where the equator $ \partial
\mathcal{B}_{\epsilon} \cap \mathcal{P}_{1} $ intersects the
half-plane containing $ v_{2} $ and bounded by the line determined by
$ v_{1} $ and $ v_{3} .$ Then we can deform the segment $ q_{1}v_{1} $
so that it curves along a great circle path $ \alpha_{1} $ from $
q_{1} $ to $ q_{2}, $ and then in a straight line path to $ v_{1} .$
See Figure ~\ref{fig:epsilon.ball}(b). Note that since the arc $
\alpha_{1} $ lies on the same plane as $ v_{2}v_{3}, $ then $
p(\alpha_{1}) \cup p(v_{2}v_{3}) $ forms a single great circle
trajectory on $ \mathcal{S} $ from $ p(v_{3}) $ to $ p(q_{2}) .$ Thus,
after this deformation, the upper bound on the total number of
crossings given above still holds.
Similarly, let $ q_{3} $ be the point of intersection between the
equator $ \partial \mathcal{B}_{\epsilon} \cap \mathcal{P}_{1} $ and
the half-plane containing $ v_{n} $ and bounded by the line determined
by $ v_{1} $ and $ v_{n-1}, $ and let $ q_{4} $ be the point at which
the edge $ v_{n}v_{1} $ intersects $ \partial \mathcal{B}_{\epsilon}
.$ Then the segment $ v_{1}q_{4} $ can be deformed so that it travels
in a straight line path from $ v_{1} $ to $ q_{3} $ and then curves
along a great circle path $ \alpha_{3} $ from $ q_{3} $ to $ q_{4}. $
See Figure ~\ref{fig:epsilon.ball}(c). As before, the arc $
\alpha_{3} $ lies on the same plane as $ v_{n-1}v_{n}, $ so $
p(v_{n-1}v_{n}) \cup p(\alpha_{3}) $ forms a single great circle
trajectory on $ \mathcal{S} $ from $ p(v_{n-1}) $ to $ p(q_{3}). $
Therefore the upper bound on the number of crossings given above still
holds after this deformation.
Finally, isotope $ P $ by moving $ v_1 $ into the interior of the
triangle $ \triangle q_{2}v_{1}q_{3} $ while curving the segments $
q_{2}v_{1} $ and $ v_{1}q_{3} $ until they coincide with an arc along
the equator $ \partial \mathcal{B}_{\epsilon} \cap \mathcal{P}_{1} ,$
as in Figure ~\ref{fig:epsilon.ball}(d). This final transformation
turns $ P $ into a non-polygonal embedding of the same (topological)
knot type; this new embedding agrees with $ P $ outside of the ball $
\mathcal{B}_{\epsilon} $ but completely avoids its interior. In the
meanwhile, the image under $ p $ of this embedding is simply a
(spherical) knot projection $ \Gamma' $ consisting of the $ n-2 $ arcs
of $ \Gamma $ (with its ends extended by $ p(\alpha_{1}) $ and $
p(\alpha_{3}) $), together with an $ (n-1) $th arc $ \alpha_{2} $
running along the equator $ \mathcal{S} \cap \mathcal{P}_{1} $ and
joining the endpoints $ p(q_2) $ and $ p(q_3) .$ Since $ \Gamma $ is
contained entirely on one side of the equator, $ \alpha_{2} $ does not
cross any other arcs. Hence the new projection has no more crossings
than it did before the last deformation, proving the following
theorem.
\begin{Thm} \label{thm:new.c.bound}
Suppose that a knot $ K $ with minimal crossing number $ c(K) $ and
minimal polygon index $ s(K) .$ Then
\begin{equation} \label{eq:crossing}
c(K) \le \frac{ \bigl(s(K) - 3 \bigr) \bigl(s(K) - 4 \bigr)}{2} .
\end{equation}
Completing the square in (\ref{eq:crossing}) shows that
\[ 2c \le s^2 - \, 7 s \, + \, 12 =
\biggl(s - \frac{7}{2}\biggr)^2 - \frac{1}{4} , \]
so that
\[ s(K) \ge \frac{7 \, + \, \sqrt{8c(K) \, + \, 1}}{2} . \]
\end{Thm}
Note that Theorem ~\ref{thm:new.c.bound} correctly predicts that the
trefoil is the only non-trivial knot which can be realized with six
edges.
\begin{figure}[tp]
\insertfig{diagram1.eps}{1.9in}
\centering{{\small \hspace{5mm} (a) \hspace{60mm} (b) }}
\insertfig{diagram2.eps}{1.9in}
\centering{{\small \hspace{5mm} (c) \hspace{60mm} (d) }}
\caption{A knot universe and several choices in over- and
under-crossings.}
\label{fig:diagrams}
\end{figure}
In the case of octagons, the new bound on crossing number becomes 10.
However, in \cite{Calvo:thesis} we systematically look at the possible
knot projections $ \Gamma' $ resulting from the deformation described
by Figure ~\ref{fig:epsilon.ball} and thereby enumerate the
topological knots which are appear in $ \mathfrak{Geo}^{(8)} .$ For example,
consider the ten-crossing knot universe shown in Figure
~\ref{fig:diagrams}(a). By appropriately choosing at each crossing
which strand goes ``over'' and which one goes ``under,'' we will
obtain a knot projection $ \Gamma' $ corresponding, as above, to some
octagon $ P .$ As we make choices in ``over'' and ``under'' crossings
we need to keep a few points in mind:
\begin{enumerate} \renewcommand {\labelenumi} {(\roman{enumi})}
\item If $ v_2 v_3 $ passes under every one of its crossings, then the
interior of the triangular disc $ \triangle v_1 v_2 v_3 $ does not
intersect the rest of $ P .$ In this case, $ P $ can be isotoped by
pushing $ v_2 $ in a straight line path to the midpoint of the line
segment $ v_1 v_3 $ until $ P $ coincides with a heptagon. A similar
isotopy exists if $ v_{7}v_{8} $ contributes only ``under'' crossings.
Therefore we need not consider these diagrams.
\item If the edges $ v_{5}v_{6} $ and $ v_{6}v_{7} $ both go under $
v_{3}v_{4}, $ as in Figure ~\ref{fig:diagrams}(b), then we can isotope
$ P $ so that the corresponding $ \Gamma' $ has two fewer crossings.
For instance, we can shrink the lengths of $ v_{5}v_{6} $ and $
v_{6}v_{7}, $ in essence performing a Reidemeister 2 move. We can
therefore ignore crossing choices which permit a reducing isotopy of
this type, delaying their analysis until we examine the resulting
reduced diagram.
\item Some choices of ``over'' and ``under'' crossings will lead to
configurations which are impossible to create with straight edges.
For instance, consider the three crossing choices made in Figure
~\ref{fig:diagrams}(c). Let $ \mathcal{P} $ be the plane containing $
v_{4}, v_{5}, $ and $ v_{6} .$ Note that the interior of edge $ v_6
v_7 $ lies entirely above the plane $ \mathcal{P}, $ since it starts
on the plane at $ v_{6} $ and then crosses over $ v_4 v_5 .$
Similarly, the interior of edge $ v_{3} v_{4} $ lies below the plane $
\mathcal{P} $ since it crosses under $ v_{5} v_{6} $ and then meets
the plane at $ v_{4} .$ This means that $ v_{3} v_{4} $ cannot cross
over $ v_6 v_7, $ as in Figure ~\ref{fig:diagrams}(c), unless one of
the two edges is bent.
\item A particularly tricky example of a bad ``over'' and ``under''
crossing choice is shown in Figure ~\ref{fig:diagrams}(d). This
diagram corresponds to an octagonal realization of the knot $ 8_{18}
,$ shown in Figure ~\ref{fig:eight.18}. Here the problem is not as
obvious as before. In fact, among all of the projections
corresponding to impossible configurations which we encounter in
\cite{Calvo:thesis}, this is the only one which is not clearly
impossible. Nonetheless, through a delicate balance between
introducing self-intersections and counting dimensions, we can show
that there is no way to construct this configuration. The details of
this argument will appear in a forthcoming paper.
\end{enumerate}
\begin{figure}[t]
\insertfig{eight18.eps}{2in}
\caption{This octagonal embedding of the knot $ 8_{18} $ cannot be
constructed with straight edges.}
\label{fig:eight.18}
\end{figure}
After considering all possible projections $ \Gamma' $ with more than
six crossings, we find that the only knots with polygon index 8 and
crossing number greater than 6 are $ 8_{19} $ and $ 8_{20} .$ Since it
is known that there are octagonal realizations of every knot $ K $
with crossing number $ c(K) \le 6, $ we obtain a complete list of the
topological knots present in $ \mathfrak{Geo}^{(8)}, $ as indicated in Theorem
~\ref{thm:classification}(iv). With the exception of $ 6_{3} , $ the
square knot $ 3_{1} - 3_{1}, $ the figure-eight knot $ 4_{1}, $ and
the unknot, every knot type in this list is chiral and therefore must
contribute at least two path-components in $ \mathfrak{Geo}^{(8)} .$ Therefore $
\mathfrak{Geo}^{(8)} $ will contain at least twenty path-components.
\section*{Acknowledgments}
I would like to thank Ken Millett, who first led me into this
wonderful subject, and who has always been happy to give me his
advice, insights, and toughest questions. I would also like to thank
Janis Cox Millett for her hospitality this summer, as the three of us
traveled through Paris, Athens, Delphi, Berlin, and Aix-en-Provence.
\bibliographystyle{amsplain}
|
\section{Introduction}
The {\sl adiabatic piston problem} is a well-known example in thermodynamics
which has given rise to continuous controversy for the last 40 years
(see \cite{gruber98}
for selected references). It was also mentioned by E. Lieb
as one of the
``problems in statistical mechanics that I would like to
see solved'' at the StatPhys 20 meeting in 1998 \cite{lieb}.
The problem is the
following. The system is a finite cylinder containing two gases separated
by an adiabatic movable piston. Initially, the piston is rigidly fixed
by a brake and the two gases are in thermal equilibrium characterized by
$\left(U_{[1]},V_{[1]},N_{[1]}\right)$
and $\left(U_{[2]},V_{[2]},N_{[2]}\right)$
(see Fig. \ref{fig1}).
At a certain time $t_0$, the brake is released
and the question is to find the final equilibrium state.
\begin{figure}
\epsfxsize=10truecm
\hspace{3.25truecm}
\epsfbox{fig1.eps}
\caption{
The adiabatic piston setting.
\label{fig1}}
\end{figure}
Within the framework of {\sl thermostatics}, one only knows that the final
equilibrium state
$\left(U_{[1]}(\infty),V_{[1]}(\infty),N_{[1]}(\infty)\right)$,
$\left(U_{[2]}(\infty),V_{[2]}(\infty),N_{[2]}(\infty)\right)$,
corresponds
to a maximum of the entropy function
$S(U,V)=S_{[1]}+S_{[2]}=S_{[1]}(U,V)+S_{[2]}(U_0-U,V-V_0)$, where
$U_0=U_{[1]}+U_{[2]}$, $V_0=V_{[1]}+V_{[2]}$, submitted
to the constraints $S_{[1]}\geq S_{[1]}(t_0)$, $S_{[2]}\geq S_{[2]}(t_0)$.
The solution of this mathematical problem yields the result that the final
pressures must necessarily be equal $p_{[1]}(\infty)=p_{[2]}(\infty)$
({\sl i.e.} mechanical
equilibrium), but nothing can be said about the final temperatures
$T_{[1]}(\infty)$,
$T_{[2]}(\infty)$ (originally Kubo had arrived at the wrong conclusion
$p_{[1]}(\infty)/T_{[1]}(\infty)=p_{[2]}(\infty)/T_{[2]}(\infty)$
\cite{kubo}). In other words the laws of
thermostatics are not sufficient to predict the final equilibrium state. Here
started the controversy, because some physicists could not
accept this limitation of thermostatics. Callen writes in \cite{callen}
``the movable adiabatic wall presents a unique problem with subtleties''.
The problem is however neither very subtle,
nor unique since it appears every time
movable adiabatic walls are present, and they are indeed essential in the
axiomatic formulation of thermostatics (see {\sl e.g.} \cite{lieb1}).
Recently, with the help of an oversimplified macroscopical model for the gases,
it was shown that the adiabatic piston problem, can be solved within the
framework of {\sl thermodynamics}\cite{gruber98}.
Equations for the time evolution were derived from the first and second law.
They show that the equilibrium state depends in an essential manner on the
viscosity of the gases and one has to solve explicitly the time
evolution to find the final state,
which is however uniquely defined. In particular in the final state the
temperatures $T_{[1]}(\infty)$ and $T_{[2]}(\infty)$ are different.
Using arguments from the {\sl kinetic theory of gases}, similar equations were
previously obtained by B. Crosignani et al. \cite{crosignani}, without the use
of the laws of thermodynamics, but neglecting the effects of fluctuations.
The conclusion were however the same as the one obtained from thermodynamics,
namely $T_{[1]}(\infty)\ne T_{[2]}(\infty)$.
At this point the controversy started again:
it was argued that the stochastic motion of the piston, induced by the
stochastic forces exerted by the particles of the gases, might lead to a
``true'' equilibrium state where $T_{[1]}(\infty)= T_{[2]}(\infty)$
and $p_{[1]}(\infty)=p_{[2]}(\infty)$.
In \cite{crosignani} it was stated ``this stochastic motion leads to
a violation of the second law'' and ``this would be equivalent to permit
heat to flow although the piston itself remains an insulator''.
It was also suggested that this evolution toward the ``true''
equilibrium state, if it exists, might be so slow that it can never be
observed in practice.
The work done by J. L. Lebowitz in 1959 \cite{lebowitz} should have given
some insight into these questions already a long time ago. Indeed, using
a model of non-interacting particles (in one dimension) to describe the gases,
J. L. Lebowitz computed for the first time in \cite{lebowitz}, the heat
conductivity of a movable ``adiabatic'' piston at lowest order in $\epsilon=
\sqrt{m/M}$ where $m$ is the mass of the gas particles and $M$ is the mass
of the piston. However, we should remark that the model discussed in
\cite{lebowitz} was slightly different than the one considered here
since the piston was
constrained to remain in mechanical equilibrium by means of an external force.
This same microscopical model, but without external force, was reintroduced by
Piasecki and Gruber in \cite{gruber1}. Starting with the
{\sl Boltzmann equation}, they obtained the exact stationary solution for
the {\sl infinite} one dimensional system, with $p_{[1]}(t)=
p_{[1]}=p_{[2]}=p_{[2]}(t)$,
$T_{[1]}(t)=T_{[1]}\ne T_{[2]}=T_{[2]}(t)$,
in the very special case $M=m$. The main conclusion
in \cite{gruber1} was that the space asymmetry ($T_{[1]}\ne T_{[2]}$)
conspires
with the stochastic motion to give a stationary non-equilibrium state with
non-zero average velocity of the ``piston'' toward the region of the
higher temperature. The same model with $M=m$ was then considered by
Piasecki and Sinai\cite{sinai} to study the finite system. They were able
to show that the piston evolves toward an equilibrium position corresponding
to uniform density and uniform distribution of velocities (thus uniform
``temperature''), for the particles throughout the system.
Simultaneously Gruber and Piasecki investigated the more physical situation
$m\ll M$ \cite{gruber2}.
Assuming that the solution of the Boltzmann equation has an
asymptotic expansion in $\epsilon=\sqrt{m/M}$, the stationary solution of the
{\sl infinite} system was computed at the order $\epsilon^2$. It was again
established that the temperature difference ($T_{[1]}\ne T_{[2]}$)
induces a
macroscopic motion toward the high temperature region despite the absence of
any macroscopical force ($p_{[1]}=p_{[2]}$). In particular, it was shown
that in this stationary state the piston behaves as a randomly moving particle
with kinetic energy $k_B\sqrt{T_{[1]}T_{[2]}}/2$,
and average velocity ${\sqrt{m}\over M}\sqrt{{\pi k_B\over 8}}
\left(\sqrt{T_{[2]}}-\sqrt{T_{[1]}}\right)$.
In all these recent articles, the physical problems of heat conduction,
viscous force, work power, as well as the approach toward the stationary,
or the equilibrium, state were not considered. These
questions are investigated in the present article.
Although we did not solve the problem mentioned by E. Lieb, which would need
a rigorous mechanical many-body approach, the following analysis answers
some of the conjectures discussed above and should contribute to a better
understanding of the physical mechanisms involved in the time evolution
of the adiabatic piston. It raises also some conjectures that one might
be able to prove.
The main conclusion of this paper is that the phenomenological macroscopic
thermodynamics, the microscopical Boltzmann equation, and the numerical
simulations, give results which co\"\i ncide with a very high degree
of accuracy when $m\ll M$. In other words there is no violation of the second
law and no paradoxes involved.
The results of this investigation show that a wall which is ``adiabatic''
when rigidly fixed (no heat transfer), becomes ``conducting'' as soon as it
is allowed to have a stochastic motion, and this is independent of the fact
that its macroscopic velocity is zero or non-zero. This means that in this
model heat transfer is associated with the stochastic motion of the piston,
while work power is related to the macroscopic (or average) velocity of the
piston $\left\langle V\right\rangle_t$.
The time evolution obtained from thermodynamics
(where the above conductivity is taken into account),
from Boltzmann equation, and from
the numerical simulations, shows that the system evolves toward a
final state where the temperatures, as well as the pressures, are equal.
However, and this is the fundamental point,
when $m\ll M$ the time scale needed
to reach this final state is so enormous (about the age of the universe
for realistic numbers), that no reasonable person would call such a
piston a conductor. More precisely our numerical analysis indicates
that as soon as $m/M$ is small enough,
the time evolution proceeds in two
stages with time scales which are very different.
In the first stage the
evolution is adiabatic and proceeds rather rapidly, with or without
oscillations, towards a state of mechanical equilibrium where the pressures
become equal, but
the temperatures remain different (this is the ``quasi-equilibrium state''
suggested in \cite{crosignani}). After this quasi-equilibrium situation is
reached the piston will slowly drift into the high temperature domain and
the system evolves, on a time scale several orders of magnitude
larger, with equal pressures and slowly varying temperatures toward
the final equilibrium state where temperatures, densities and pressures are
equal on both sides of the piston.
On the other hand for
$m/M\simeq 1$, the adiabatic piston is a good conductor and thermal
equilibrium $T_{[1]}(\infty)= T_{[2]}(\infty)$ is reached rapidly.
Notice that on a very short time scale, there exists a stage where
the piston behaves as if the system is infinite. In this stage, if the initial
pressures are different, the piston is accelerated while if the pressures are
initially equal the piston evolves with a (small) constant velocity inside the
high temperature fluid.
In the following analysis the microscopical model is a one dimensional
system of point particles which do not interact except for purely
elastic collisions. It is exactly this model which has been simulated on
a computer.
For the analytical analysis we introduce three main assumptions.
First we assume that the correlations between the velocities
of the fluid particles and
the velocity of the piston can be neglected.
Then we assume that
the velocity distribution functions for the fluid particles on
both faces of the piston admit a scaling function with respect to their
mass $m$, as it is the case with Maxwellian distributions.
However, because of the simulations (Sec.\ref{simul}), and the
exact solution of Piasecki and Sinai \cite{sinai},
we do not assume Maxwellian
distribution for the velocity of the fluid particles. The temperatures
$T_{[1]}(t)$ and $T_{[2]}(t)$
on both faces of the piston are then defined by the
second moment of the velocity of the fluid particles.
Finally we assume that the moments $\left\langle V^s\right\rangle_t$ of the velocity
$V$ of the piston, $s\geq 1$,
have an asymptotic expansion in powers of $\epsilon=\sqrt{m/M}$.
To discuss the solutions of the time evolution equations, we shall
either assume that the fluids densities and temperatures
are constant (in the case of
an infinite system), or uniform (in the case of a finite system).
The microscopical model
is defined in Sec.\ref{micro}. In Sec.\ref{conserve} we
use the conservation laws of the problem to derive the dynamical
equations for the
first and second moment of the piston velocity distribution and give in
Sec.\ref{arbitrary} these equations for arbitrary moments $\left\langle V^s\right\rangle_t$.
The relation between the microscopical model and a phenomenological
thermodynamic approach is discussed in Sec.\ref{pheno} where we derive
a microscopical definition of the friction coefficients
and the heat conductivity. The analysis
of the stationary non-equilibrium, and the equilibrium, states is presented in
Sec.\ref{stationary}, while the time evolution is discussed in
Sec.\ref{timeevol}.
The numerical simulations, the comparison with the dynamic equations, and the
thermodynamic equations, is the subject of Sec.\ref{simul}. Finally,
conclusions are given in Sec.\ref{conclusion}.
\section{Microscopical model}\label{micro}
We consider the fluids to be perfect gases made of $N_{[1]}$, respectively
$N_{[2]}$,
identical non-interacting
point particles of mass $m$, making purely elastic collisions, and
contained in a fixed cylinder of length $L$. The basis of the cylinder is a
circle with area $A$. The adiabatic piston is a rigid solid ({\sl i.e.}
no internal degree of freedom), with mass $M$, described by the microscopical
variables $(X,V={dX\over dt})$ (see Fig. \ref{fig1}).
We assume that the cylinder is an ideal constraint, which has no physical
properties: the collisions of the particles on the cylinder are purely
elastic, and there is no friction between the piston and the cylinder.
For $t\leq 0$, the piston is fixed $(X=X_0,V=0)$ and both fluids are in
thermal equilibrium, with temperatures $T_{[1]}$,
respectively $T_{[2]}$,
described by the Maxwellian distribution functions:
\begin{eqnarray}
&& \rho_{[1]}(x,v;t\leq 0)=\rho_{[1]}\phi_{T_{[1]}}(v)\quad{\rm for}
\quad 0\leq x< X_0
\nonumber\\
&& \rho_{[2]}(x,v;t\leq 0)=\rho_{[2]}\phi_{T_{[2]}}(v)\quad{\rm for}
\quad X_0< x\leq L
\label{1}
\end{eqnarray}
where
\begin{equation}
\rho_{[1]}={1\over A}{N_{[1]}\over X_0}={1\over A}n_{[1]},\quad
\rho_{[2]}={1\over A}{N_{[2]}\over L-X_0}={1\over A}n_{[2]}
\end{equation}
and
\begin{equation}
\phi_{T}(v)=
\sqrt{{m\over 2\pi k_BT}}\exp\left(-{mv^2\over 2k_BT}\right).
\label{max}
\end{equation}
At time $t=0$, the break is released and the piston can move freely. It will
thus have a stochastic motion induced by collisions with the fluids
and $(X,V)$ become random variables.
The problem is
to investigate the random motion of the piston.
To simplify the discussion, we consider the system to be one-dimensional.
In this case the adiabatic piston is just a point particle with mass $M$.
For $t>0$, the piston moves freely. Since the collisions are assumed to be
purely elastic, then if the masses $M$ and $m$ have velocities $V$ and $v$
before the collision, their velocities after the collision are given by
\begin{eqnarray}
&& v'=v-{2M\over M+m}(v-V)\nonumber\\
&& V'=V+{2m\over M+m}(v-V)
\label{coll}
\end{eqnarray}
The main assumption of our analysis is that the time evolution is such
that it is possible to neglect the correlations between the velocity of
the piston and the velocities
of the fluids particles. This means that we assume
that the joint probability distribution to find, at time $t$,
the piston at position $X$ with velocity $V$ and a fluid particle at position
$X\pm |\epsilon|$ with velocity $v$ can be expressed by
\begin{eqnarray}
\Phi_{[1]}(X,V,X,v;t)&=&\Phi(X,V;t)\lim_{\epsilon\to 0}\rho_{[1]}(X-|\epsilon|,v;t)=
\Phi(X,V;t)n_{[1]}(t)\phi_{[1]}(v,t)\nonumber\\
\Phi_{[2]}(X,V,X,v;t)&=&\Phi(X,V;t)\lim_{\epsilon\to 0}\rho_{[2]}(X+|\epsilon|,v;t)=
\Phi(X,V;t)n_{[2]}(t)\phi_{[2]}(v,t)
\label{assump}
\end{eqnarray}
where we have introduced the time dependent
particle densities on the faces of the
piston $n_{[i]}$ and the normalized fluids velocity
distribution functions on the faces of the piston $\phi_{[i]}(v,t)$
for $i=1,2$.
In the case of an infinite cylinder, we shall consider that
$n_{[1]}(t)$ and $n_{[2]}(t)$ are constant, equal to
the fluid initial densities,
and that $\phi_{[i]}(v,t)=\phi_{T_{[i]}}(v)$ ($i=1,2$).
This assumption is justified if $m\ll M$, since
in this case there is a vanishing
probability that the piston interact back with the perturbation it causes in
the state of the surrounding fluids (no recollision). In other words, the
piston always ``sees'' on both sides the unperturbed initial states.
For the finite system, when $m\ll M$, we expect the average velocity
of the piston to be small, and therefore the time needed for a macroscopical
but small displacement of the piston will be much larger than the relaxation
time of the fluids
(if the initial pressures difference is not too large).
We shall then assume in Sec.\ref{arbitrary}
that at all time $t$, the
fluids on both side of the piston are homogeneous and thus that
the distribution functions $\rho_{[1]}(x,v,t)$, $0<x<X$, and
$\rho_{[2]}(x,v,t)$, $X<x<L$, are independent from the
position $x$. We then have
\begin{equation}
n_{[1]}(t)={N_{[1]}\over \left\langle X\right\rangle_t}, \quad
n_{[2]}(t)={N_{[2]}\over L-\left\langle X\right\rangle_t}
\end{equation}
and the time-dependent temperatures $T_{[1]}(t)$ and $T_{[2]}(t)$ are
defined by the second moment of the fluid
velocity.
For the following analysis, we introduce the small parameters
\begin{equation}
\alpha={2m\over M+m}\quad {\rm and}\quad \epsilon=\sqrt{{m\over M}}.
\end{equation}
\section{Conservation laws}\label{conserve}
We use the conservation laws present in the problem to determine the dynamical
equations
for moments of the piston velocity distribution.
In particular, the conservation
of momentum and energy leads to equations for the first and second moment of
the piston velocity distribution.
\subsection{Linear momentum}
Let $\Pi_{[1]}(t)$ and $\Pi_{[2]}(t)$ denote the
linear momentum of the fluids on
the left and on the right of the piston at time $t$. Using the collision
equations (\ref{coll}) and the assumption (\ref{assump}) the force
exerted by the left fluid on the piston is
\begin{equation}
F^{[1]\to p}=-\lim_{\delta t\to 0}
\left\langle {\delta \Pi_{[1]}(t)\over \delta t}\right\rangle=
M\alpha\int_{-\infty}^\infty dV \Phi_\epsilon(V;t)\int_V^\infty dv\,
n_{[1]}(t)\phi_{[1]}(v,t)(v-V)^2
\label{8}\end{equation}
where $\Phi_\epsilon(V;t)$ is the time dependent
velocity distribution function of the piston.
Similarly the force exerted by the right fluid on the piston is
\begin{equation}
F^{[2]\to p}=
-\lim_{\delta t\to 0}\left\langle {\delta \Pi_{[2]}(t)\over \delta t}
\right\rangle=
-M\alpha\int_{-\infty}^\infty dV \Phi_\epsilon(V;t)\int_{-\infty}^V dv\,
n_{[2]}(t)\phi_{[2]}(v,t)(v-V)^2.
\label{9}
\end{equation}
Therefore, the conservation law of linear momentum
\begin{equation}
M{d\over dt}\left\langle V\right\rangle_t-F^{[1]\to p}-F^{[2]\to p}=
M{d\over dt}\left\langle V\right\rangle_t+ \lim_{\delta t\to 0}\left[
\left\langle {\delta \Pi_{[1]}(t)\over \delta t}
\right\rangle+\left\langle {\delta \Pi_{[2]}(t)\over \delta t}
\right\rangle\right]=0
\label{10}
\end{equation}
can be written in the following form
\begin{eqnarray}
{d\over dt}\left\langle V\right\rangle_t &=& \alpha
\int_{-\infty}^\infty dV \Phi_\epsilon(V;t)\int_0^\infty dv\,
\left[n_{[1]}(t)\phi_{[1]}(v,t)(v-V)^2-n_{[2]}(t)\phi_{[2]}(-v,t)(v+V)^2
\right] \nonumber\\
&& \quad -\alpha \int_{-\infty}^\infty dV \Phi_\epsilon(V;t)\int_0^V dv\,(v-V)^2
\left[n_{[1]}(t)\phi_{[1]}(v,t)+n_{[2]}(t)\phi_{[2]}(v,t)\right].
\label{mom0}
\end{eqnarray}
Using the expansion
\begin{equation}
\phi_{[i]}(v,t)=\sum_{k=0}^\infty
{v^k\over k!}\phi_{[i]}^{\prime(k)}(t),\quad (i=1,2)
\label{expansion}
\end{equation}
where
\begin{equation}
\phi^{\prime(k)}_{[i]}(t)
= \left.\left({d\over dv}\right)^k\phi_{[i]}(v,t)\right|_{v=0},
\quad (i=1,2),
\end{equation}
we obtain
\begin{eqnarray}
{d\over dt}\left\langle V\right\rangle_t &=& \alpha\left[n_{[1]}(t)\left\langle v^2\right\rangle^{[1]}_t-
n_{[2]}(t)\left\langle v^2\right\rangle^{[2]}_t\right]\nonumber\\
&&\quad-\left\langle V\right\rangle_t 2\alpha \left[n_{[1]}(t)\left\langle v\right\rangle^{[1]}_t+
n_{[2]}(t)\left\langle v\right\rangle^{[2]}_t\right]\nonumber \\
&&\quad\quad
+\left\langle V^2\right\rangle_t \alpha \left[n_{[1]}(t)\left\langle
v^0\right\rangle^{[1]}_t
-n_{[2]}(t)\left\langle v^0\right\rangle^{[2]}_t\right]\nonumber\\
&&\quad\quad\quad
-\sum_{k=0}^\infty \left\langle V^{k+3}\right\rangle_t{2\alpha\over (k+3)!}
\left[n_{[1]}(t)\phi_{[1]}^{\prime(k)}(t)
+n_{[2]}(t)\phi_{[2]}^{\prime(k)}(t)\right]
\label{mom1}
\end{eqnarray}
where for any non negative integers $s$
\begin{eqnarray}
\left\langle v^s\right\rangle^{[1]}_t
&=& \int_0^\infty dv\,\phi_{[1]}(v,t)v^s,
\nonumber\\
\left\langle v^s\right\rangle^{[2]}_t
&=& \int_0^\infty dv\,\phi_{[2]}(-v,t)v^s,\quad {\rm and}
\nonumber\\
\left\langle V^s\right\rangle_t &=&
\int_{-\infty}^{\infty} dV\, \Phi_\epsilon(V;t)V^s.
\label{phi0}
\end{eqnarray}
We should insist on the fact that in the definitions of
$\left\langle v^s\right\rangle^{[i]}_t$ the integration is on $v$ positive.
In particular $\left\langle v^0\right\rangle^{[i]}_t$ is not equal to 1.
We should note that the evolution equation (\ref{mom1}) for
the average velocity of the
piston depends on higher moments of the piston velocity.
We should also remark that if the term
$n_{[1]}(t)\left\langle v^2\right\rangle^{[1]}_t-n_{[2]}(t)\left\langle v^2\right\rangle^{[2]}_t$ is zero
(which means, as we shall see in Sec.\ref{pheno}, that the
pressures exerted by the fluids are equal) and
if we neglect the fluctuations of the piston velocity
($\left\langle V^s\right\rangle_t=\left\langle V\right\rangle_t^s$) we find the stable equilibrium solution $\left\langle V\right\rangle_t=0$.
The effects of the fluctuations is the main problem to be investigated and
will give a very different picture.
\subsection{Energy}
Similarly, let $E_{[1]}(t)$ and $E_{[2]}(t)$ denote the energy
of the fluids on
the left and on the right of the piston at time $t$.
It is equal to the kinetic energy, since the particles do not interact
except for elastic collisions. Using the collision equations (\ref{coll})
and the assumption (\ref{assump}),
we have
\begin{equation}
\lim_{\delta t\to 0}\left\langle
{\delta E_{[1]}(t)\over \delta t}\right\rangle=
-M\alpha\int_{-\infty}^\infty dV \Phi_\epsilon(V;t)\int_V^\infty dv\,
n_{[1]}(t)\phi_{[1]}(v,t)(v-V)^2
\left[V+{\alpha\over 2}(v-V)\right]
\label{16}
\end{equation}
and
\begin{equation}
\lim_{\delta t\to 0}\left\langle
{\delta E_{[2]}(t)\over \delta t}\right\rangle=
M\alpha\int_{-\infty}^\infty dV \Phi_\epsilon(V;t)\int_{-\infty}^V dv\,
n_{[2]}(t)\phi_{[2]}(v,t)(v-V)^2
\left[V+{\alpha\over 2}(v-V)\right].
\end{equation}
Therefore, the conservation of energy
\begin{equation}
{1\over 2}M{d\over dt}\left\langle V^2\right\rangle_t+
\lim_{\delta t\to 0}\left[
\left\langle {\delta E_{[1]}(t)\over \delta t}\right\rangle+
\left\langle {\delta E_{[2]}(t)\over \delta t}\right\rangle\right]=0
\label{17}
\end{equation}
leads to
\begin{eqnarray}
&&{d\over dt}\left\langle V^2\right\rangle_t =
2\alpha
\int_{-\infty}^\infty dV \Phi_\epsilon(V;t)\int_0^\infty dv\, \nonumber\\
&&\quad\quad
\left\{n_{[1]}(t)\phi_{[1]}(v,t)(v-V)^2\left[V+{\alpha\over 2}(v-V)\right]
-n_{[2]}(t)\phi_{[2]}(-v,t)(v+V)^2\left[V-{\alpha\over 2}(v+V)\right]
\right\} \nonumber\\
&& \quad -2\alpha \int_{-\infty}^\infty dV
\Phi_\epsilon(V;t)\int_0^V dv\,(v-V)^2\left[V+
{\alpha\over 2}(v-V)\right]
\left[n_{[1]}(t)\phi_{[1]}(v,t)+n_{[2]}(t)\phi_{[2]}(v,t)\right].
\end{eqnarray}
With the expansion (\ref{expansion}) and the definitions (\ref{phi0})
this last expression can be written as
\begin{eqnarray}
{d\over dt}\left\langle V^2\right\rangle_t &=& \alpha^2\left[n_{[1]}(t)
\left\langle v^3\right\rangle^{[1]}_t+
n_{[2]}(t)\left\langle v^3\right\rangle^{[2]}_t\right]\nonumber\\
&& +\left\langle V\right\rangle_t\alpha(2-3\alpha)\left[n_{[1]}(t)\left\langle v^2\right\rangle^{[1]}_t-
n_{[2]}(t)\left\langle v^2\right\rangle^{[2]}_t\right]\nonumber\\
&& -\left\langle V^2\right\rangle_t\alpha(4-3\alpha)\left[n_{[1]}(t)\left\langle v\right\rangle^{[1]}_t+
n_{[2]}(t)\left\langle v\right\rangle^{[2]}_t\right]\nonumber\\
&& +\left\langle V^3\right\rangle_t\alpha(2-\alpha)\left[n_{[1]}(t)\left\langle v^0\right\rangle^{[1]}_t-
n_{[2]}(t)\left\langle v^0\right\rangle^{[2]}_t\right]\nonumber\\
&& -\sum_{k=0}^\infty \left\langle V^{k+4}\right\rangle_t{2\alpha\over (k+4)!}
(2k+8-3\alpha)\left[n_{[1]}(t)\phi_{[1]}^{\prime(k)}(t)
+n_{[2]}(t)\phi_{[2]}^{\prime(k)}(t)\right]
\label{mom2}
\end{eqnarray}
We thus obtained an equation for the evolution of the second moment of the
piston velocity.
Let us remark that even in case of mechanical equilibrium where
$n_{[1]}(t)\left\langle v^2\right\rangle^{[1]}_t-
n_{[2]}(t)\left\langle v^2\right\rangle^{[2]}_t=0$
the initial state $\left\langle V^s\right\rangle_0=0$
is unstable if
$n_{[1]}(t)\left\langle v^3\right\rangle^{[1]}_t
+n_{[2]}(t)\left\langle v^3\right\rangle^{[2]}_t\ne 0$.
In conclusion,
if we do not neglect the correlations in (\ref{mom1}), we see that
even in case of mechanical equilibrium,
the initial state $\left\langle V^s\right\rangle_0=0$ is unstable
and thus the piston will start to move.
\section{Equation for arbitrary moment}\label{arbitrary}
In the previous section, we derived equations for the first and second
moments of the piston velocity which are coupled with higher order moments.
We derive in this section
the infinite set of coupled equations for all the moments
of the piston velocity.
We use the Boltzmann's equation for the piston velocity
distribution derived in \cite{gruber1} under the assumption (\ref{assump})
\begin{eqnarray}
{\partial \over \partial t}\Phi_\epsilon(V;t)&=&
\int_{-\infty}^\infty dv\, |V-v|\left[
\Theta(V-v) n_{[1]}(t)\phi_{[1]}\left(v-(2-\alpha)(v-V);t\right)
\Phi_\epsilon\left(V+\alpha(v-V);t\right)\right.\nonumber\\
&& \quad\left.
+\Theta(v-V)n_{[2]}(t)\phi_{[2]}\left(v-(2-\alpha)(v-V);t\right)
\Phi_\epsilon\left(V+\alpha(v-V);t\right)\right.\nonumber\\
&& \quad\left.-\Theta(v-V)n_{[1]}(t)\phi_{[1]}(v,t)\Phi_\epsilon(V,t)
-\Theta(V-v)n_{[2]}(t)\phi_{[2]}(v,t)\Phi_\epsilon(V,t)\right]
\end{eqnarray}
to obtain the equations for the velocity moments
\begin{eqnarray}
{d\over dt}\left\langle V^s\right\rangle_t &=& \int_{-\infty}^\infty dV V^s {\partial \over \partial t}
\Phi_\epsilon(V;t)
\nonumber\\
&=& \left\langle \int_V^\infty dv\, n_{[1]}(t)\phi_{[1]}(v,t)(v-V)\left\{\left[
V(1-\alpha)+\alpha v\right]^s-V^s\right\}\right\rangle_t\nonumber\\
&& \quad
-\left\langle \int_{-\infty}^V dv\, n_{[2]}(t)\phi_{[2]}(v,t)(v-V)\left\{\left[
V(1-\alpha)+\alpha v\right]^s-V^s\right\}\right\rangle_t
\end{eqnarray}
where $\left\langle A(V)\right\rangle_t=\int_{-\infty}^\infty dV A(V)
\Phi_\epsilon(V;t)$.
This equation can be written in the following form
\begin{eqnarray}
&&{d\over dt}\left\langle V^s\right\rangle_t =
\left\langle \int_0^\infty dv\, n_{[1]}(t)\phi_{[1]}(v,t)(v-V)\left\{\left[
V(1-\alpha)+\alpha v\right]^s-V^s\right\}\right\rangle_t\nonumber\\
&& \quad\quad +\left\langle \int_0^\infty dv\, n_{[2]}(t)
\phi_{[2]}(-v,t)(v+V)\left\{\left[
V(1-\alpha)-\alpha v\right]^s-V^s\right\}\right\rangle_t\nonumber\\
&& \quad\quad -\left\langle \int_0^V dv\, (v-V)\left\{\left[
V(1-\alpha)+\alpha v\right]^s-V^s\right\}
\left[ n_{[1]}(t)\phi_{[1]}(v,t)+
n_{[2]}(t)\phi_{[2]}(v,t)\right]\right\rangle_t.
\end{eqnarray}
Using the expansions (\ref{expansion}) and
\begin{equation}
\left[V+\alpha (v-V)\right]^s-V^s=\sum_{q=1}^{s}{s!\over q!(s-q)!}
\alpha^q(v-V)^qV^{s-q},
\end{equation}
we obtain
\begin{equation}
{d\over dt}\left\langle V^s\right\rangle_t=\sum_{r=0}^\infty A_{s,r}(t)\left\langle V^r\right\rangle_t
\label{serie0}
\end{equation}
where one has the explicit expression of $A_{s,r}(t)$ as a function of
the parameter $\alpha$. In the following, it will be more convenient to have
the explicit expression of $A_{s,r}(t)$ as a function of the parameter
$\epsilon=\sqrt{m/M}$.
Motivated by the Maxwellian distribution (\ref{max}), we
introduce the scaled velocity $w=\sqrt{m}v$
and the corresponding normalized distribution functions
$f_{[i]}(w,t)$
\begin{equation}
\phi_{[i]}(v,t)=\sqrt{m}f_{[i]}(\sqrt{m} v,t),\quad \int_{-\infty}^\infty dw\,
f_{[i]}(w,t)=1,\quad (i=1,2).
\label{scale}
\end{equation}
We then have from Eqs.~(\ref{phi0},\ref{scale}),
\begin{equation}
\left\langle v^q\right\rangle_t^{[i]}=m^{-q/2}\int_0^\infty dw f_{[i]}(w;t)
w^q= m^{-q/2} \left\langle w^q\right\rangle_t^{[i]},\,\,(i=1,2)
\end{equation}
and
\begin{equation}
\phi_{[i]}^{\prime(k)}(t)=m^{(k+1)/2}f_{[i]}^{\prime(k)}(t).
\end{equation}
For any integer $r$ we introduce the functions $K_r^{[i]}(t)$ and
$K_r(t)$ defined in the following manner.
For any $q\geq 0$, and $i=1,2$,
\begin{eqnarray}
K_{-2+q}^{[i]}(t)&=& M^{-q/2}n_{[i]}(t)\left\langle
w^{q}\right\rangle_t^{[i]},\\
K_{-3-q}^{[i]}(t)&=& M^{(q+1)/2}n_{[i]}(t) f_{[i]}^{\prime(q)}(t),\\
K_r(t)&=& K_r^{[1]}(t)-(-1)^r K_r^{[2]}(t)\quad{\rm for}\,\, r\geq -2,
\label{kr1}\\
K_r(t)&=& K_r^{[1]}(t)+K_r^{[2]}(t)\quad{\rm for}\,\, r\leq -3.
\label{kr2}
\end{eqnarray}
With these definitions, we have
\begin{equation}
{d\over dt}\left\langle V^s\right\rangle_t=\sum_{r=0}^\infty A_{s,r}(t)\left\langle V^r\right\rangle_t
\label{serie}
\end{equation}
where the coefficients $A_{s,r}(t)$ are given by
\begin{equation}
A_{s,r}(t)={\epsilon^{|s-r-1|}\over (1+\epsilon^2)^{s}}P_{s,r}(\epsilon^2) K_{s-r-1}(t).
\label{34}
\end{equation}
The functions $K_r(t)$, Eqs.~(\ref{kr1},\ref{kr2}),
depend only on the densities
and the velocity distribution functions of the fluid particles
on both faces of the piston, while $P_{s,r}(\epsilon^2)$ is a polynomial in
$\epsilon^2$, of order $\min(r,s-1)$, with constant coefficients, given by
\begin{equation}
P_{s,r}(\epsilon^2)=
(1-\epsilon^2)^{r-1}\left[(s+1-r)
-\epsilon^2(s+1+r)\right] {2^{s-r} s!\over r! (s-r+1)!}, \quad r\leq s-1;
\end{equation}
\begin{eqnarray}
P_{s,s}(\epsilon^2)&=& {(1-\epsilon^2)^{s-1}\left[1-\epsilon^2(2s+1)
\right]-(1+\epsilon^2)^s\over \epsilon^2};\\
P_{s,s+1}(\epsilon^2)&=& {(1+\epsilon^2)^{s}-(1-\epsilon^2)^s
\over \epsilon^2};
\end{eqnarray}
\begin{equation}
P_{s,s+2+k}(\epsilon^2) = -2\sum_{q=0}^{s-1}(-1)^q (2\epsilon^2)^q(1+\epsilon^2)^{s-q-1}
{s!(q+2)\over (s-1-q)!(k+q+3)!}\quad r\geq s+2.
\end{equation}
One can verify that Eq.(\ref{serie}) for $s=1,2$ are identical to the
equations for the first and the second moment $(\ref{mom1},\ref{mom2})$
derived in the preceding section.
\section{From kinetic theory to thermodynamics:
Friction coefficient and heat conduction}\label{pheno}
In this section we relate the microscopical model to a phenomenological
thermodynamic approach. This will suggest
microscopical definitions for ``pressure'', ``viscous force'', ``temperature''
and ``heat flux''.
Using a very primitive model for the fluids, the equations for the time
evolution of the piston problem were derived in \cite{gruber98}
from the first and second laws of the thermodynamics. They read
\begin{eqnarray}
(M+\delta M) {dV\over dt} &=& A(p_{[1]}-p_{[2]})
-(\lambda_{[1]}+\lambda_{[2]})V;\quad V={dX\over dt}
\label{eqv}\\
{dS_{[1]}\over dt}&=&{\lambda_{[1]}\over T_{[1]}}V^2
+{\kappa\over T_{[1]}}(T_{[2]}-T_{[1]})
\label{eqs1}\\
{dS_{[2]}\over dt}&=&{\lambda_{[2]}\over T_{[2]}}V^2
-{\kappa\over T_{[2]}}(T_{[2]}-T_{[1]})
\label{eqs2}\end{eqnarray}
where $X$ and $V$ are the macroscopic position and
velocity of the piston of mass $M$,
$\delta MV^2/2$ represents the kinetic energy of the fluids, $A$ is the area
of the cylinder,
$p_{[1]},p_{[2]},\lambda_{[1]},\lambda_{[2]},\kappa$ are phenomenological
time-dependent functions of the variables
$(X,V,S_{[1]},S_{[2]})$ and denote the pressures, the friction coefficients,
and the heat conductivity.
$S_{[1]}$ and $S_{[2]}$ denote the entropy of the fluids on
the left and on the right
of the piston.
For ideal fluids in one dimension, the thermal equation (\ref{eqs1}) and
(\ref{eqs2}) are equivalent to
\begin{eqnarray}
{1\over 2} N_{[1]} k_B {dT_{[1]}\over dt} &=& \lambda_{[1]} V^2
-{N_{[1]} k_B T_{[1]}\over X}V
+\kappa(T_{[2]}-T_{[1]}) \label{42}\\
{1\over 2} N_{[2]} k_B {dT_{[2]}\over dt} &=& \lambda_{[2]} V^2
+{N_{[2]} k_B T_{[2]}\over L-X}V
-\kappa(T_{[2]}-T_{[1]})\label{43}.
\end{eqnarray}
In thermodynamics, the adiabatic piston is defined by
$\kappa=0$. One should remark that in this case
($\kappa=0$),
if the velocity of the piston is sufficiently small so that the term in $V^2$
can be neglected, the evolution of the piston is damped, but
reversible in the thermodynamical sense, characterized by the adiabats
\begin{eqnarray}
S_{[1]}(t) = S_{[1]} \quad &{\rm and}& \quad S_{[2]}(t) = S_{[2]}\nonumber\\
{\sl i.e.}\quad\quad\quad
T_{[1]}X^2 = {\rm cte} \quad && \quad T_{[2]}(L-X)^2 = {\rm cte}.
\end{eqnarray}
In particular, the solution $p_{[1]}=p_{[2]}$, $V=0$ is stable.
To make the connection with
the thermodynamical equations (\ref{eqv}), we write the microscopic
equation for conservation of momentum (\ref{mom0}) or (\ref{mom1}) in the
form
\begin{eqnarray}
&& M\left(1+{m\over M}\right){d\over dt}\left\langle V\right\rangle_t =
2m\left[n_{[1]}(t)\left\langle v^2\right\rangle^{[1]}_t-
n_{[2]}(t)\left\langle v^2\right\rangle^{[2]}_t\right]\nonumber \\
&&\quad -\left\langle V\right\rangle_t 2m n_{[1]}(t) \left[2\left\langle v\right\rangle^{[1]}_t
-\left\langle V\right\rangle_t\left\langle v^0\right\rangle^{[1]}_t
+{1\over \left\langle V\right\rangle_t}\int_0^{\left\langle V\right\rangle_t} dv\,\phi_{[1]}(v,t)(v-\left\langle V\right\rangle_t)^2\right]\nonumber \\
&&\quad -\left\langle V\right\rangle_t 2m n_{[2]}(t) \left[2\left\langle v\right\rangle^{[2]}_t
+\left\langle V\right\rangle_t\left\langle v^0\right\rangle^{[2]}_t
+{1\over \left\langle V\right\rangle_t}\int_0^{\left\langle V\right\rangle_t} dv\,\phi_{[2]}(v,t)(v-\left\langle V\right\rangle_t)^2\right]\nonumber \\
&&\quad +\left[\left\langle V^2\right\rangle_t
-\left\langle V\right\rangle_t^2\right]
2m \left[n_{[1]}(t)\left\langle v^0\right\rangle^{[1]}_t-
n_{[2]}(t)\left\langle v^0\right\rangle^{[2]}_t\right]\nonumber \\
&&\quad -\sum_{k=0}^\infty \left[\left\langle V^{k+3}\right\rangle_t-
\left\langle V\right\rangle_t^{k+3}\right]{4m\over (k+3)!}
\left[n_{[1]}(t)\phi_{[1]}^{\prime(k)}(t)
+n_{[2]}(t)\phi_{[2]}^{\prime(k)}(t)\right].
\label{mom1bis}
\end{eqnarray}
We are thus led to identify $\left\langle V\right\rangle_t$ with the macroscopic velocity $V(t)$
of the piston, and to define the ``pressure'', the ``temperature'', and the
``friction coefficient'', on both sides ($i=1,2$) of the
piston by
\begin{equation}
p_{[i]}(t)=n_{[i]}(t)k_BT_{[i]}(t)=2mn_{[i]}(t)
\left\langle v^2\right\rangle^{[i]}_t=2n_{[i]}(t)
\left\langle w^2\right\rangle^{[i]}_t=2MK_0^{[i]}(t)
\label{tempdefine}
\end{equation}
\begin{equation}
\lambda_{[i]}(V,t)=2mn_{[i]}(t)\left[2\left\langle v\right\rangle^{[i]}_t
+(-1)^i V\left\langle v^0\right\rangle^{[i]}_t
+{1\over V}\int_0^V dv\,\phi_{[i]}(v,t)(v-V)^2\right]
\label{defcoef}
\end{equation}
which gives in particular
\begin{equation}
\lambda_{[i]}(V=0)=4\sqrt{m}n_{[i]}\left\langle w\right\rangle^{[i]}
=4\epsilon MK_{-1}^{[i]},
\label{lllv0}
\end{equation}
\begin{equation}
\lambda_{[1]}(V)+\lambda_{[2]}(V)=M\left[4\epsilon K_{-1}
-2\epsilon^2K_{-2}V+{\cal O}(\epsilon^3 V^2))
\right],
\end{equation}
and shows that the condition $\lambda_{[i]}(V)\geq 0$ will be satisfied in the
domain considered here where the average piston velocity is
much smaller than the average absolute velocity of a fluid particle,
$V\ll\left\langle v\right\rangle^{[i]}_t$. Notice that the temperature
definition Eq.(\ref{tempdefine}) is consistent with a Maxwellian definition
of the temperature.
We thus have from Eq.(\ref{mom1bis},\ref{8},\ref{9},\ref{10})
\begin{equation}
M(1+\epsilon^2){d\over dt}\left\langle V\right\rangle_t =\left(p_{[1]}(t)-p_{[2]}(t)\right)
-\left(\lambda_{[1]}(V,t)+\lambda_{[2]}(V,t)\right)\left\langle V\right\rangle_t+F_{\rm st}(t),
\label{fst}
\end{equation}
and
\begin{eqnarray}
F^{[1]\to p}(t)&=&-F^{p\to [1]}(t)={1\over 1+\epsilon^2}\left[p_{[1]}(t)
-\lambda_{[1]}(V,t)\left\langle V\right\rangle_t
+F_{\rm st}^{[1]}\right]\label{51}\\
F^{[2]\to p}(t)&=&-F^{p\to [2]}(t)={1\over 1+\epsilon^2}\left[-p_{[2]}(t)
-\lambda_{[2]}(V,t)\left\langle V\right\rangle_t
+F_{\rm st}^{[2]}\right]
\end{eqnarray}
where the stochastic forces are
\begin{eqnarray}
F_{\rm st}^{[1]}&=&\left[\left\langle V^2\right\rangle_t-\left\langle V\right
\rangle_t^2\right]
2m n_{[1]}(t)\left\langle v^0\right\rangle^{[1]}_t-
\sum_{k=0}^\infty \left[\left\langle V^{k+3}\right\rangle_t-
\left\langle V\right\rangle_t^{k+3}\right]{4m\over (k+3)!}
n_{[1]}(t)\phi_{[1]}^{\prime(k)}(t)\nonumber\\
F_{\rm st}^{[2]}&=&-\left[\left\langle V^2\right\rangle_t-\left\langle V\right
\rangle_t^2\right]
2m n_{[2]}(t)\left\langle v^0\right\rangle^{[2]}_t-
\sum_{k=0}^\infty \left[\left\langle V^{k+3}\right\rangle_t-
\left\langle V\right\rangle_t^{k+3}\right]{4m\over (k+3)!}
n_{[2]}(t)\phi_{[2]}^{\prime(k)}(t).
\end{eqnarray}
and $F_{\rm st}=F_{\rm st}^{[1]}+F_{\rm st}^{[2]}$.
We now compute the heat flux from one side of the piston to the
other side in the microscopic model. This will give us a
microscopic definition of the heat conductivity $\kappa$. We will see
that the microscopical equations are compatible with the
phenomenological thermodynamical equations in which we take into account
a heat flux characterized by a conductivity $\kappa\ne 0$
which comes from the
fluctuations of the microscopical model.
From the first law of thermodynamics
\begin{equation}
{d\over dt}E_{[1]}=P_W^{p\to [1]}+P_Q^{[2]\to [1]}
\end{equation}
where $P_W^{p\to [1]}$, respectively $P_Q^{[2]\to [1]}$,
denote the power transmitted
to the fluid on the left in the form of work, resp. in the form of heat,
together with the thermodynamic definition of the work power, {\sl i.e.}
\begin{equation}
P_W^{p\to [1]}(t)=F^{p\to [1]}(t)V(t)
\end{equation}
we are led to define, through Eqs.(\ref{51},\ref{16},\ref{17},\ref{serie}),
the microscopic heat power, or heat flux, and the
microscopic heat conductivity $\kappa$
\begin{eqnarray}
P_Q^{[2]\to [1]}(t)&=&\left\langle {dE_{[1]}\over dt}\right\rangle
-F^{p\to [1]}(t)\left\langle V\right\rangle_t\nonumber\\
&=& -{1\over 2}MA_{2,0}^{[1]}+M\sum_{r=0}^\infty \left[A_{1,r}^{[1]}
\left\langle V^r\right\rangle_t\left\langle V\right\rangle_t-{1\over 2}A_{2,r+1}^{[1]}\left\langle
V^{r+1}\right\rangle_t
\right]\nonumber\\
&=& M{4\epsilon^2\over 1+\epsilon^2}K_0^{[1]}\left\langle V\right\rangle_t-M{4\epsilon\over (1+\epsilon^2)^2}
\left\{\left[(1+\epsilon^2)\left\langle V\right\rangle_t^2-(1-{\epsilon^2\over 2})\left\langle V^2\right\rangle_t\right]K_{-1}^{[1]}
+{1\over 2}K_{1}^{[1]}\right\}\nonumber \\
&& \quad +M\sum_{r=2}^\infty \left[A_{1,r}^{[1]}\left\langle
V^r\right\rangle_t\left\langle V\right\rangle_t
-{1\over 2}A_{2,r+1}^{[1]}\left\langle V^{r+1}\right\rangle_t\right]\nonumber\\
&=&\kappa (T_{[2]}-T_{[1]})
\label{heatflux}\end{eqnarray}
where $A_{2,r}^{[i]}$ are defined by Eq.(\ref{34}) with $K_{1-r}^{[i]}$.
At order ${\cal O}(\epsilon)$, it reduces to
\begin{eqnarray}
P_Q^{[2]\to [1]}&=&
4\epsilon MK_{-1}^{[1]}\left[\left\langle V^2\right\rangle_t-\left\langle V\right\rangle_t^2\right]-2\epsilon MK_1^{[1]}+{\cal O}(\epsilon^2)
\nonumber\\
&=&
4{\sqrt{m}\over M}n_{[1]}(t)\left\{M\left[\left\langle V^2\right\rangle_t-\left\langle V\right\rangle_t^2\right]
\left\langle w\right\rangle_t^{[1]}-{1\over 2}\left\langle
w^3\right\rangle_t^{[1]}\right\}
+{\cal O}(\epsilon^2).
\label{pq}
\end{eqnarray}
To compare our definition (\ref{heatflux}) with the heat conductivity
obtained in \cite{lebowitz}, let us consider the case of the
{\sl infinite} system. In this case the coefficients $K_r^{[i]}$ and the
densities $n_{[i]}$ are independent on time.
As we shall see in Sec.\ref{stationary}, for any stationary state of the
infinite system we have
\begin{equation}
\left\langle V^2\right\rangle_\infty-\left\langle V\right\rangle^2_\infty
={K_1\over 2 K_{-1}}+{\cal O}(\epsilon)
\end{equation}
which thus gives for a stationary state
\begin{equation}
P_Q^{[2]\to [1]}={2\sqrt{m}\over M}n_{[1]}n_{[2]}
{\left\langle w^3\right\rangle^{[1]}\left\langle w\right\rangle^{[2]}
-\left\langle w^3\right\rangle^{[2]}\left\langle w\right\rangle^{[1]}\over
\left\langle w^2\right\rangle^{[1]}\left\langle w\right\rangle^{[2]}
+\left\langle w^2\right\rangle^{[2]}\left\langle w\right\rangle^{[1]}}+{\cal O}(\epsilon^2).
\end{equation}
In particular, for a Maxwellian distribution of the fluid particles,
this last equation reads
\begin{equation}
P_Q^{[2]\to [1]}={k_B\over M}\sqrt{{8m k_B \over \pi}}n_{[1]}n_{[2]}
{\sqrt{T_{[1]}T_{[2]}}\over n_{[1]}\sqrt{T_{[1]}}+n_{[2]}\sqrt{T_{[2]}}}
(T_{[2]}-T_{[1]})
+{\cal O}(\epsilon^2).
\end{equation}
Therefore the heat conductivity is
\begin{equation}
\kappa={k_B\over M}\sqrt{{8mk_B\over \pi}}n_{[1]}n_{[2]}
{\sqrt{T_{[1]}T_{[2]}}\over n_{[1]}\sqrt{T_{[1]}}+n_{[2]}\sqrt{T_{[2]}}}
+{\cal O}(\epsilon^2).
\label{conductivity}
\end{equation}
Similarly, for the infinite system the friction coefficient
(\ref{defcoef}) can be written as
\begin{equation}
\lambda_{[1]}(V)=n_{[1]}\sqrt{{8k_B T_{[1]} m\over \pi }}-n_{[1]}mV
+n_{[1]}m\sqrt{{2m\over \pi k_B T_{[1]}}}{1\over V}\int_0^V dv\,
{\rm exp}\left(-{mv^2\over 2k_B T_{[1]}}\right)v^2
\label{friction}
\end{equation}
It is interesting to note that for a stationary state of the {\sl infinite}
system and Maxwellian distribution for the fluids,
the expression for the
friction coefficient $\lambda(V=0)$ Eq.~(\ref{friction}), and for the
heat conductivity $\kappa$ Eq.~(\ref{conductivity}) are, to lowest order in
$\epsilon$,
identical with those first derived by J..L. Lebowitz for a slightly
different model \cite{lebowitz}. In \cite{lebowitz} the piston is
constrained in a state of mechanical equilibrium ({\sl i.e.}
$\left\langle V^k\right\rangle=0$ for all odd integer $k$) by mean of
an external potential and the heat flux is just given by $dE_{[1]}/dt$.
In our model there is no external potential and
in the stationary state the piston has a constant
non-zero average velocity; we thus have to take into account the work power
in the definition of the heat flux Eq.(\ref{heatflux}).
We should also remark that in
\cite{lebowitz} the friction coefficient is introduced via the
Ornstein-Uhlenbeck process describing the motion of the piston,
while our definition is associated
with the mechanical equation of motion Eq.(\ref{eqv}) and thus the factor
$M^{-1}$ difference between $\lambda$ in \cite{lebowitz} and our $\lambda(V=0)$,
Eq.(\ref{friction}). Finally, we want to
stress that in the definition (\ref{defcoef})
and (\ref{heatflux}) of $\lambda$ and $\kappa$, the
fluid particles can have any arbitrary distribution and the system is not
necessarily in a stationary state.
To conclude this section we remark that we can use
the thermodynamic equation
\begin{equation}
{dE_{[1]}\over dt}= T_{[1]}{dS_{[1]}\over dt} -p_{[1]}{dX\over dt}
+\delta M {dX\over dt}
{d^2X\over dt^2}
\end{equation}
to define the time evolution of the microscopic entropy by
\begin{eqnarray}
T_{[1]}{dS_{[1]}\over dt}&=&
\left\langle {dE_{[1]}\over dt}\right\rangle +p_{[1]}\left\langle V\right\rangle_t-m \left\langle V\right\rangle_t{d\over dt}\left\langle V\right\rangle_t
\nonumber \\
&=& -{M\over 2}\sum_{k=0}^\infty A^{[1]}_{2,r}\left\langle V^r\right\rangle_t
+p_{[1]}\left\langle V\right\rangle_t-m \left\langle V\right\rangle_t{d\over dt}\left\langle V\right\rangle_t\nonumber \\
&=& {2\sqrt{m}\over M}n_{[1]}\left[2\left\langle w\right\rangle_t^{[1]}M\left\langle V^2\right\rangle_t-
\left\langle w^3\right\rangle_t^{[1]}\right]+{\cal O}(\epsilon^2)
\end{eqnarray}
{\sl i.e.} with (\ref{defcoef}) and (\ref{pq})
\begin{eqnarray}
T_{[1]}{dS_{[1]}\over dt}&=& \lambda_{[1]}\left\langle V^2\right\rangle_t-2{\sqrt{m}\over M}n_{[1]}(t)
\left\langle w^3\right\rangle_t^{[1]}+{\cal O}(\epsilon^2)\nonumber \\
&=& \lambda_{[1]}\left\langle V\right\rangle_t^2+{\lambda_{[1]}\over M}\left[M(\left\langle V^2\right\rangle_t-\left\langle V\right\rangle_t^2)-
{\left\langle w^3\right\rangle_t^{[1]}\over 2\left\langle
w\right\rangle_t^{[1]}} \right]
+{\cal O}(\epsilon^2)\nonumber\\
&=& \lambda_{[1]}\left\langle V\right\rangle_t^2+ P_Q^{2\to 1},
\end{eqnarray}
which is consistent with the thermodynamical equation (\ref{eqs1}).
Again, assuming a Maxwellian distribution for
the fluid particles, we will have
\begin{equation}
T_{[1]}{dS_{[1]}\over dt}=
\lambda_{[1]}\left\langle V\right\rangle_t^2+{\lambda_{[1]}\over M}\left[M(\left\langle V^2\right\rangle_t-\left\langle V\right\rangle_t^2)-k_BT_{[1]}\right]
+{\cal O}(\epsilon^2).
\label{68}
\end{equation}
In summary, if initially $\left\langle V^r\right\rangle_{t=0}=0$ for all
$r\geq 1$, then
as long as the contributions
\begin{equation}
{\lambda_{[i]}\over M}\left[M(\left\langle V^2\right\rangle_t-\left\langle V\right\rangle_t^2)-k_BT_{[i]}\right], \quad (i=1,2),
\end{equation}
as well as the stochastic force $F_{\rm st}$, can be neglected,
the time evolution Eqs.(\ref{fst},\ref{68})
will co\"\i ncide with the thermodynamic evolution of
the adiabatic piston Eq.(\ref{eqv}-\ref{eqs2}) with
$\kappa=0$.
This will be the case as long as
$\left\langle V^r\right\rangle_t\simeq \left\langle V\right\rangle_t^r$, and $\left\langle V\right\rangle_t\gg
k_B\max(T_{[1]},T_{[2]})/M$.
Furthermore, as long as $\left\langle V\right\rangle_t^2\ll1$ the evolution is described by the adiabats
$S_{[1]}(t)=S_{[1]}$, $S_{[2]}(t)=S_{[2]}$. Therefore during
this first stage the piston
evolves according to (\ref{eqv})
towards a state of mechanical equilibrium, {\sl i.e.} $p_{[1]}\simeq p_{[2]}$,
but not thermal equilibrium $(T_{[1]}\neq T_{[2]})$. As soon as
the pressures are
approximately equal, the above assumptions are no longer valid, and
as we shall see in Sec.\ref{timeevol} the
system will evolve toward a state of thermal equilibrium.
Numerical simulations will be presented in Sec.\ref{simul}
to show the time scales on which these evolutions take place.
\section{Stationary non-equilibrium and equilibrium state}\label{stationary}
\subsection{Preliminary remarks}\label{prelimi}
We have seen in Sec.\ref{arbitrary}
that the time evolution of the piston is
characterized by the equations
\begin{equation}
{d\over dt}\left\langle V^s\right\rangle_t=C_s+\sum_{r=1}^\infty A_{s,r}\left\langle V^r\right\rangle_t,\quad s\geq 1
\label{evol}
\end{equation}
where $C_s(t)=A_{s,0}(t)$ and for any $r\geq 0$
\begin{equation}
A_{s,r}(t)={\epsilon^{|s-r-1|}\over (1+\epsilon^2)^s}P_{s,r}(\epsilon^2)K_{s-r-1}(t).
\end{equation}
In particular for $s=1$ and $s=2$ we have
(where $\stackrel{*}{=}$ denotes ``for Maxwellian
distributions''),
\begin{eqnarray}
{d\over dt}\left\langle V\right\rangle_t &=& {1\over M(1+\epsilon^2)}\left[(p_{[1]}-p_{[2]})
-\lambda\left\langle V\right\rangle_t+m(n_{[1]}-n_{[2]})\left\langle V^2\right\rangle_t+\cdots\right]
\label{701}\\
{d\over dt}\left\langle V^2\right\rangle_t &=& {2\over M(1+\epsilon^2)^2}\left[(1-2\epsilon^2)
(p_{[1]}-p_{[2]})\left\langle V\right\rangle_t-\lambda\left\langle V^2\right\rangle_t+2\epsilon M K_1+\cdots\right]
\label{702}
\end{eqnarray}
with
\begin{eqnarray}
p_{[1]}-p_{[2]} &=& M2K_0\stackrel{*}{=} k_B(n_{[1]}T_{[1]}-n_{[2]}T_{[2]})\nonumber\\
\lambda &=& M4\epsilon K_{-1}\stackrel{*}{=} \sqrt{m}\sqrt{{8k_B\over \pi}}
\left(n_{[1]}\sqrt{T_{[1]}}+n_{[2]}\sqrt{T_{[2]}}\right)\nonumber\\
2M\epsilon K_1 &\stackrel{*}{=} & {\sqrt{m}\over M}\sqrt{{8k_B\over \pi}}k_B
\left(n_{[1]}T_{[1]}^{3/2}+n_{[2]}T_{[2]}^{3/2}\right)\nonumber\\
2K_{-2} &\stackrel{*}{=} & n_{[1]}-n_{[2]}
\label{71a}
\end{eqnarray}
To lowest order, {\sl i.e.} in the limit $\epsilon\to 0$, Eq.(\ref{evol})
together with the definition (\ref{tempdefine}) of the pressure yields
\begin{eqnarray}
{d\over dt}\left\langle V^s\right\rangle_t &=& 2sK_0(t)\left\langle V^{s-1}\right\rangle_t\nonumber\\
&=& s{1\over M}[p_{[1]}(t)-p_{[2]}(t)]\left\langle V^{s-1}\right\rangle_t.
\end{eqnarray}
As discussed in Sec.\ref{micro},
for an infinite system with $\epsilon\ll 1$, the
assumptions
\begin{equation}
n_{[i]}(t)=n_{[i]},\quad \phi_{[i]}(v,t)=\phi_{T_{[i]}}(v),\quad(i=1,2)
\end{equation}
are justified and give $p_{[1]}(t)=p_{[1]}$, $p_{[2]}(t)=p_{[2]}$ constant.
\noindent
In this case, for the initial conditions
\begin{equation}
\left\langle V^{s}\right\rangle_{t=0}
=\left(\left\langle V\right\rangle_{t=0}
\right)^s
\end{equation}
we obtain at lowest order
\begin{eqnarray}
\left\langle V\right\rangle_t &=&\left\langle V\right\rangle_{t=0}+{1\over M}(p_{[1]}-p_{[2]})t
\nonumber\\
\left\langle V^s\right\rangle_t &=&\left(\left\langle V\right\rangle_t\right)^s
\end{eqnarray}
and thus there is no stationary state in the {\sl infinite} system
unless $p_{[1]}=p_{[2]}$ in the limit $\epsilon\to 0$.
On the other hand, for the finite system, if the evolution is sufficiently
slow and $\epsilon\ll 1$, we may assume that the following homogeneity
conditions hold:
\begin{equation}
n_{[1]}(t)={N_{[1]}\over \left\langle X\right\rangle_t},\quad n_{[2]}(t)= {N_{[2]}\over L-\left\langle X\right\rangle_t},\quad
\left\langle E_{[i]}\right\rangle_t=N_{[i]}m\left
\langle v^2\right\rangle_t^{[i]}
\equiv {1\over 2}
N_{[i]}k_BT_{[i]}(t),\quad (i=1,2)
\label{homogene}
\end{equation}
where we do not assume Maxwellian distribution of the velocities but
define the temperatures by the second moments of the fluid velocity
distribution Eq.(\ref{tempdefine}).
Under this homogeneity condition, the time evolution for
$T_{[1]}$ and $T_{[2]}$ are therefore
\begin{equation}
{dT_{[i]}(t)\over dt}={2\over N_{[i]}k_B}\left\langle
{d\over dt}E_{[i]}\right\rangle=
-{M\over N_{[i]} k_B}\sum_{r=0}^\infty A_{2,r}^{[i]}\left\langle V^r\right\rangle_t,\quad(i=1,2).
\label{evolt}\end{equation}
In the limit $\epsilon\to 0$ we have from
Eqs.(\ref{evol},\ref{homogene},\ref{evolt})
\begin{eqnarray}
M{d^2\over dt^2}\left\langle X\right\rangle_t &=& k_B\left[{N_{[1]}\over \left\langle X\right\rangle_t}T_{[1]}
-{N_{[2]}\over L-\left\langle X\right\rangle_t}T_{[2]}\right]
\nonumber\\
{dT_{[1]}\over dt} &=& -{2\over \left\langle X\right\rangle_t}T_{[1]}{d\over dt}\left\langle X\right\rangle_t \nonumber\\
{dT_{[2]}\over dt} &=& {2\over L-\left\langle X\right\rangle_t}T_{[2]}{d\over dt}\left\langle X\right\rangle_t
\end{eqnarray}
which yields the ``adiabats''
\begin{equation}
T_{[1]}(t)\left\langle X\right\rangle_t^2=\gamma_{[1]}={\rm cte};\quad T_{[2]}(t)(L-\left\langle X\right\rangle_t)^2
=\gamma_{[2]}={\rm cte}
\end{equation}
and the undamped oscillatory motion
\begin{equation}
M{d^2\over dt^2}\left\langle X\right\rangle_t=k_B\left[{N_{[1]}\gamma_{[1]}\over \left\langle X\right\rangle_t^3}-
{N_{[2]}\gamma_{[2]}\over (L-\left\langle X\right\rangle_t)^3}\right].
\end{equation}
In conclusion, to lowest order in $\epsilon$, there is no approach toward an
equilibrium state in the finite system while
a stationary state in the infinite system
can exist only if the initial pressures are
equal $(p_{[1]}=p_{[2]})$.
This conclusion is consistent with the previous result,
since for finite $M$
the condition $\epsilon\to 0$ means that $m\to 0$ and thus the
friction coefficient and the heat conductivity are both zero.
\subsection{Stationary state of the infinite system}
\label{6.2}
In this section, we discuss the stationary states of the infinite system.
In this case the densities $n_{[1]}$, $n_{[2]}$, and the coefficients
$K_r$ are constant. Moreover we consider the moments $\left\langle w^q\right\rangle$
which appear in the definition of $K_r$ to be given by the Maxwellian
distributions {\sl i.e.}
\begin{eqnarray}
&& \left\langle w^0\right\rangle^{[i]}={1\over 2},\quad
\left\langle w^1\right\rangle^{[i]}={1\over 2\pi}\sqrt{2k_BT_{[i]}}, \nonumber\\
&& \left\langle w^2\right\rangle^{[i]}={1\over 2}k_BT_{[i]},\quad
\left\langle w^3\right\rangle^{[i]}={1\over 2\pi}(2k_BT_{[i]})^{3/2},\quad (i=1,2).
\end{eqnarray}
The stationary states of the infinite system
are given by the solution of the equations
\begin{equation}
0=\sum_{r\geq 0}\epsilon^{|s-r-1|}P_{s,r}(\epsilon^2)K_{s-r-1}\left\langle V^r\right\rangle_\infty.
\label{station}
\end{equation}
From Eqs.(\ref{701},\ref{702}), it follows immediately that
\begin{eqnarray}
&& \left\langle V\right\rangle_\infty = {p_{[1]}-p_{[2]}\over \lambda}
+{m\over \lambda}(n_{[1]}-n_{[2]})\left\langle V^2\right\rangle_\infty+\cdots\label{811}\\
&& \left\langle V^2\right\rangle_\infty\left[1-(1-2\epsilon^2){p_{[1]}-p_{[2]}\over \lambda}
{m\over \lambda}(n_{[1]}-n_{[2]})\right]=(1-2\epsilon^2)
\left({p_{[1]}-p_{[2]}\over \lambda}\right)^2 +2{\epsilon\over\lambda}MK_1+\cdots
\label{812}
\end{eqnarray}
{\sl i.e.} at lowest order in $\epsilon$ we have
\begin{eqnarray}
\left\langle V\right\rangle_\infty &=& {p_{[1]}-p_{[2]}\over \lambda}+
\left({p_{[1]}-p_{[2]}\over \lambda}\right)^2{m\over \lambda}{1\over k_B}
\left({p_{[1]}\over T_{[1]}}-{p_{[2]}\over T_{[2]}}\right)+{m\over M}
{p_{[1]}T_{[2]}-p_{[2]}T_{[1]}\over \lambda \sqrt{T_{[1]}T_{[2]}}}\label{813}\\
\left\langle V^2\right\rangle_\infty &=& \left({p_{[1]}-p_{[2]}\over \lambda}\right)^2+
{k_B\over M}\sqrt{T_{[1]}T_{[2]}}\label{814}
\end{eqnarray}
which shows that is impossible to have a stationary state with
$\left\langle V\right\rangle_\infty=0$, $p_{[1]}=p_{[2]}$, $T_{[1]}\ne T_{[2]}$.
In the following we shall assume that the solution of Eq.(\ref{station}) has an
asymptotic expansion in $\epsilon$
\begin{equation}
\left\langle V^r\right\rangle_\infty=\sum_{l=0}^\infty \epsilon^l \left\langle V^r\right\rangle_\infty^{(l)}.
\label{assumption2}
\end{equation}
Then in the limit $\epsilon\to 0$ we must have
\begin{equation}
K_0=0,\quad {\sl i.e.}\quad p_{[1]}=p_{[2]}
\label{presseq}
\end{equation}
which can also be seen explicitly with Eqs.(\ref{811},\ref{812}) since
$\lambda={\cal O}(\epsilon)$.
To take into account the possibility of non-zero pressure difference for
the infinite system, and to have results valid for finite systems where
the equilibrium densities might depend on $\epsilon$, we now consider that
the densities $n_{[1]}$, $n_{[2]}$ are function of $\epsilon$
\begin{equation}
n_{[i]}(\epsilon)=\sum_{l=0}^\infty \epsilon^l n_{[i]}^{(l)}\quad (i=1,2)
\end{equation}
which then implies that the coefficients $K_r$ are also functions of $\epsilon$
\begin{equation}
K_r(\epsilon)=\sum_{l=0}^\infty \epsilon^l K_r^{(l)}
\end{equation}
The condition (\ref{presseq}) for a stationary solution is now
\begin{equation}
K_0^{(0)}=\lim_{\epsilon\to 0} K_0(\epsilon)=0
\end{equation}
{\sl i.e.}
\begin{equation}
n_{[1]}^{(0)}T_{[1]}-n_{[2]}^{(0)}T_{[2]}=0
\end{equation}
or
\begin{equation}
\lim_{\epsilon\to 0}[p_{[1]}(\epsilon)-p_{[2]}(\epsilon)]=p_{[1]}^{(0)}-p_{[2]}^{(0)}=0.
\end{equation}
Therefore Eq.(\ref{station}) can be simplified by $\epsilon$ to give
\begin{eqnarray}
0&=& P_{s,s-1}\left({K_0\over \epsilon}\right)\left\langle V^{s-1}\right\rangle_\infty+\nonumber\\
& &\quad + \sum_{r=0}^{s-2}\epsilon^r \left[P_{s,s-2-r}K_{1+r}
\left\langle V^{s-2-r}\right\rangle_\infty
+P_{s,s+r}K_{-1-r}\left\langle V^{s+r}\right\rangle_\infty\right]\nonumber\\
& &\quad + \sum_{r=0}^{\infty}\epsilon^{s-1+r}P_{s,2s-1+r}K_{-s-r}
\left\langle V^{2s-1+r}\right\rangle_\infty.
\label{station2}
\end{eqnarray}
We then solve (\ref{station2}) at successive order in $\epsilon$.
At order $\epsilon^0$, we have
\begin{eqnarray}
0&=& 2K_0^{(1)}-4K_{-1}^{(0)}\left\langle V\right\rangle_\infty^{(0)}\quad(s=1)\nonumber\\
0&=& 4K_0^{(1)}\left\langle V\right\rangle_\infty^{(0)}+4K_1^{(0)}-8K_{-1}^{(0)}
\left\langle V^{2}\right\rangle_\infty^{(0)}\quad(s=2)\nonumber\\
0&=& 2sK_0^{(1)}\left\langle V^{s-1}\right\rangle_\infty^{(0)}+2s(s-1)K_1^{(0)}
\left\langle V^{s-2}\right\rangle_\infty^{(0)}
-4sK_{-1}^{(0)}\left\langle V^{s}\right\rangle_\infty^{(0)}\quad(s\geq 3)
\end{eqnarray}
which yields
\begin{eqnarray}
\left\langle V\right\rangle_\infty^{(0)} &=& {K_0^{(1)}\over 2 K_{-1}^{(0)}}=V_0
={1\over 2\sqrt{M}}
{n_{[1]}^{(1)}\left\langle w^2 \right\rangle^{[1]}-n_{[2]}^{(1)}\left\langle w^2 \right\rangle^{[2]}\over
n_{[1]}^{(0)}\left\langle w \right\rangle^{[1]}+n_{[2]}^{(0)}\left\langle w \right\rangle^{[2]}}\nonumber\\
&=& \sqrt{{\pi k_B \over 8M}T_{[1]}T_{[2]}}\lim_{\epsilon\to 0}{1\over \epsilon}
{p_{[1]}(\epsilon)-p_{[2]}(\epsilon)\over p_{[1]}(\epsilon)\sqrt{T_{[2]}}
+p_{[2]}(\epsilon)\sqrt{T_{[1]}}},
\label{91}
\end{eqnarray}
\begin{eqnarray}
\left\langle V^2\right\rangle_\infty^{(0)}-\left(\left\langle V\right\rangle_\infty^{(0)}\right)^2 &=&
{K_1^{(0)}\over 2 K_{-1}^{(0)}}=\Delta={1\over 2M}
\lim_{\epsilon\to 0}{n_{[1]}(\epsilon)\left\langle w^3 \right\rangle^{[1]}
+n_{[2]}(\epsilon)\left\langle w^3 \right\rangle^{[2]}\over
n_{[1]}(\epsilon)\left\langle w \right\rangle^{[1]}+n_{[2]}(\epsilon)\left\langle w \right\rangle^{[2]}}\nonumber\\
&=& {k_B\over M}\sqrt{T_{[1]}T_{[2]}}
\lim_{\epsilon\to 0}
{p_{[1]}(\epsilon)\sqrt{T_{[1]}}+p_{[2]}(\epsilon)\sqrt{T_{[2]}}
\over p_{[1]}(\epsilon)\sqrt{T_{[2]}}
+p_{[2]}(\epsilon)\sqrt{T_{[1]}}},
\label{v2}
\end{eqnarray}
and
\begin{equation}
\left\langle V^s\right\rangle_\infty^{(0)}
=\sum_{k=0}^{[s/2]}{s!\over 2^k k!(s-2k)!}V_0^{s-2k}s^k.
\end{equation}
Therefore at order $\epsilon^0$, the distribution function for the velocity of the
piston is
\begin{equation}
\Phi^{(0)}(V)={1\over \sqrt{2\pi \Delta}}
\exp\left[-{(V-V_0)^2\over 2\Delta}\right].
\end{equation}
Let us remark that we recover Eqs.(\ref{813},\ref{814}) at order
zero.
At order $\epsilon^1$, introducing the notation
\begin{equation}
k_r^{(l)}={K_r^{(l)}\over 2K_{-1}^{(0)}};\quad k_0^{(0)}=0;
\quad k_0^{(1)}=V_0;\quad k_1^{(0)}=\Delta
\end{equation}
Eq.(\ref{station2}) yields
\begin{eqnarray}
\left\langle V\right\rangle_\infty^{(1)} &=& -2k_{-1}^{(1)}V_0+k_{0}^{(2)}+k_{-2}^{(0)}
(\Delta+V_0^2)\label{102}\\
\left\langle V^2\right\rangle_\infty^{(1)} &=& V_0\left\langle V\right\rangle_\infty^{(1)}-2k_{-1}^{(1)}
(\Delta+V_0^2)
+k_{0}^{(2)}V_0+k_1^{(1)}+k_{-2}^{(0)}V_0(3\Delta+V_0^2)\\
\left\langle V^s\right\rangle_\infty^{(1)} &=& V_0\left\langle V^{s-1}\right\rangle_\infty^{(1)}+(s-1)\Delta\left\langle
V^{s-2}\right\rangle_\infty^{(1)}-2k_{-1}^{(1)}\left\langle V^{s}\right\rangle_\infty^{(0)}
+k_{0}^{(2)}\left\langle V^{s-1}\right\rangle_\infty^{(0)}\nonumber\\
&& \quad+(s-1)k_{1}^{(1)}\left\langle V^{s-2}\right\rangle_\infty^{(0)}
+{2\over 3}(s-1)(s-2)k_{2}^{(0)}\left\langle V^{s-3}\right\rangle_\infty^{(0)}
+k_{-2}^{(0)}\left\langle V^{s+1}\right\rangle_\infty^{(0)}.
\end{eqnarray}
As one can verify the distribution function at order $\epsilon^1$ is
\begin{eqnarray}
\Phi_\epsilon(V) &=&{1\over \sqrt{2\pi\Delta}}\exp\left[-{1\over 2\Delta}
\left(V-\sqrt{{\pi k_B \over 8M}T_{[1]}T_{[2]}}
{p_{[1]}(\epsilon)-p_{[2]}(\epsilon)\over p_{[1]}(\epsilon)\sqrt{T_{[2]}}
-p_{[2]}(\epsilon)\sqrt{T_{[1]}}}\right)^2\right]
\nonumber\\
&&\quad\times \left\{ 1+\epsilon\left[a_1(V-V_0)+a_2(V^2-\Delta-V_0^2)
+a_3(V^3-3\Delta V_0 -V_0^3)\right]\right\}
\label{101}
\end{eqnarray}
where $V_0$, $\Delta$ are given by Eqs.(\ref{91},\ref{v2}) and
\begin{eqnarray}
a_1\Delta^3 &=&k_{-1}^{(1)}2V_0\Delta^2-k_{1}^{(1)}V_0\Delta
-{2\over 3} k_{2}^{(0)}(\Delta-V_0^2)\\
a_2\Delta^3 &=& -k_{-1}^{(1)}\Delta^2+{1\over 2}k_{1}^{(1)}\Delta
-{2\over 3} k_{2}^{(0)}V_0\\
3a_3\Delta^3 &=&k_{-2}^{(0)}\Delta^2+{2\over 3} k_{2}^{(0)}.
\end{eqnarray}
Let us remark that for $p_{[1]}=p_{[2]}$ then $V_0=0$ and
we recover the results of Gruber and
Piasecki \cite{gruber1}.
One can then successively obtain higher order terms for the velocity
moments.
\noindent In conclusion, for the infinite system we observe a stationary state
with non-zero average
velocity $\left\langle V\right\rangle_\infty$,
which depends on $n_{[1]}$, $n_{[2]}$, $T_{[1]}$,
$T_{[2]}$,
given at order $\epsilon$ by Eqs.(\ref{813},\ref{102}).
It is interesting to note that this result present some analogy and
differences with those obtained by J.L. Lebowitz in \cite{lebowitz}, where
he considered an adiabatic piston constrained around the origin by
an external force: The coefficient $\Delta$ Eq.(\ref{v2}) is exactly
the same as in \cite{lebowitz}; however for
$p_{[1]}-p_{[2]}={\cal O}(\epsilon)\ne 0$ we have a drift
$V_0$, while in \cite{lebowitz} the external force give
$\left\langle V^{2k+1}\right\rangle_\infty^{(0)}=0$ ($k\geq 0$).
To conclude the discussion on the stationary states of the infinite system,
we investigate under what condition there exists a stationary
state with zero average velocity {\sl i.e.} $\left\langle V\right\rangle_\infty=0$.
From Eq.(\ref{811}) it follows immediately that
\begin{equation}
p_{[1]}-p_{[2]}=-m(n_{[1]}-n_{[2]})\left\langle V^2\right\rangle_\infty+{\cal O}(\epsilon^3).
\end{equation}
However one has to check that a stationary solution of Eq.(\ref{station}) does
exist with $\left\langle V\right\rangle_\infty=0$ at all order in $\epsilon$.
From Eqs.(\ref{91},\ref{102}), $\left\langle V\right\rangle_\infty=0$ up to order $\epsilon$ iff
\begin{equation}
K_0^{(0)} = K_0^{(1)}=0,\quad{\sl i.e.}\quad
\lim_{\epsilon\to 0}(p_{[1]}(\epsilon)-p_{[2]}(\epsilon))/\epsilon=0
\label{equal1}
\end{equation}
and
\begin{equation}
K_0^{(2)} + K_{-2}^{(0)}\Delta =0.
\end{equation}
Expressed differently, $\left\langle V\right\rangle_\infty=0$ up to order $\epsilon$ iff
\begin{eqnarray}
&& n_{[1]}^{(0)}T_{[1]}-n_{[2]}^{(0)}T_{[2]}=0\nonumber\\
&& n_{[1]}^{(1)}T_{[1]}-n_{[2]}^{(1)}T_{[2]}=0\nonumber\\
&& n_{[1]}^{(2)}T_{[1]}-n_{[2]}^{(2)}T_{[2]}=-{M\over k_B}\Delta
\left[n_{[1]}^{(0)}-n_{[2]}^{(0)}\right].
\label{equal2}
\end{eqnarray}
This last equation is equivalent to
\begin{equation}
p_{[1]}-p_{[2]}=-m\left[n_{[1]}^{(0)}
-n_{[2]}^{(0)}\right]\left\langle V^2\right\rangle_\infty^{(0)}
\end{equation}
and, in this case
\begin{eqnarray}
\left\langle V^2\right\rangle_\infty^{(0)} &=& \Delta={k_B\over M}\sqrt{T_{[1]}T_{[2]}}\\
\left\langle V^2\right\rangle_\infty^{(1)} &=& k_1^{(1)}-2k_{-1}^{(1)}\Delta.
\end{eqnarray}
At order $\epsilon^2$, we have (taking $s=1$)
\begin{equation}
\left\langle V\right\rangle_\infty^{(2)}=k_0^{(3)}-2k_{-1}^{(1)}
\left\langle V\right\rangle_\infty^{(1)}-2k_{-1}^{(2)}V_0
+k_{-2}^{(1)}\left\langle V^2\right\rangle_\infty^{(0)}+k_{-2}^{(0)}\left\langle V^2\right\rangle_\infty^{(1)}
-{1\over 3}k_{-3}^{(0)}\left\langle V^3\right\rangle_\infty^{(0)}.
\end{equation}
Therefore $\left\langle V\right\rangle_\infty=0$ up to order $\epsilon^2$ iff
\begin{equation}
K_0^{(3)}+K_{-2}^{(1)}\Delta +K_{-2}^{(0)}(k_{1}^{(1)}-2k_{-1}^{(1)}\Delta)=0
\end{equation}
Finally at order $\epsilon^q$, we have (taking $s=1$)
\begin{equation}
0=2[K_0]^{(q+1)}-4[K_{-1}\left\langle V\right\rangle_\infty]^{(q)}
+2[K_{-2}\left\langle V^2\right\rangle_\infty]^{(q-1)}
-\sum_{r\geq 3}{4\over r!}[K_{-r}\left\langle V_r\right\rangle_\infty]^{(q+1-r)}.
\end{equation}
Therefore $\left\langle V\right\rangle_\infty=0$ up to order $\epsilon^q$ iff
all $K_0^{(l)}$ are uniquely determined by the functions obtained
at previous order.
In conclusion, under the assumption Eq.(\ref{assumption2})
and given the density
$n_{[2]}$, and the distributions $\phi_{[1]}(v)$, $\phi_{[2]}(v)$, with
$T_{[1]}\ne T_{[2]}$, then the average velocity of the piston in the
stationary state is
non-zero for all densities $n_{[1]}$
except for one special value $n_{[1]}=n_{[1]}(\epsilon)$
for which $\left\langle V\right\rangle_\infty=0$.
For this special solution, one has
\begin{eqnarray}
n_{[1]}^{(0)}\left\langle w^2\right\rangle^{[1]} &=& n_{[2]}\left\langle w^2\right\rangle^{[2]}\\
n_{[1]}^{(1)} &=& 0\\
n_{[1]}^{(2)}\left\langle w^2\right\rangle^{[1]} &=& -[n_{[1]}^{(0)}\left\langle w^0\right\rangle^{[1]}-
n_{[2]}\left\langle w^0\right\rangle^{[2]}]\Delta\\
{1\over M}n_{[1]}^{(q)}\left\langle w^2\right\rangle^{[1]} &=&
-[K_{-2}\left\langle V^2\right\rangle_\infty]^{(q-2)}
+\sum_{r\geq 3}{2\over r!}[K_{-r}\left\langle V_r\right\rangle_\infty]^{(q-r)}\quad (q\geq 3).
\end{eqnarray}
In this state we have (see Sec.\ref{pheno})
\begin{equation}
P_Q^{2\to 1}=-{1\over \sqrt{M}}2\epsilon n_{[2]} \left\langle w^2\right\rangle^{[2]}
{\left\langle w^3\right\rangle^{[1]}\left\langle w\right\rangle^{[2]}
-\left\langle w^3\right\rangle^{[2]}\left\langle w\right\rangle^{[1]}\over
\left\langle w^2\right\rangle^{[1]}\left\langle w\right\rangle^{[2]}
+\left\langle w^2\right\rangle^{[2]}\left\langle w\right\rangle^{[1]}}
+{\cal O}(\epsilon^2)
\end{equation}
and thus there is a constant heat flux for any distribution $\phi_{[1]}(v)$,
$\phi_{[2]}(v)$ such that $\left\langle w^3\right\rangle^{[1]}\left\langle w\right\rangle^{[2]}\ne
\left\langle w^3\right\rangle^{[2]}\left\langle w\right\rangle^{[1]}$.
For Maxwellian distribution, this means that there is a constant flux
for any $T_{[1]}\ne T_{[2]}$.
\noindent Therefore there exists a special non-equilibrium stationary state with
zero average velocity of the piston, which is however distinct
from the one given in \cite{lebowitz} since in our case
$\left\langle V^{2k+1}\right\rangle_\infty$ is not zero (for $k\geq 1$).
\subsection{Equilibrium state of the finite system}
\label{equilfinite}
Any equilibrium state of the finite system is necessarily a stationary
state with $\left\langle V\right\rangle_\infty=0$ and $P_Q^{[2]\to [1]}=0$ where, for
$\left\langle V\right\rangle_\infty =0$ (see Eq.(\ref{heatflux})),
\begin{eqnarray}
P_Q^{[2]\to [1]}=-{2M\over (1+\epsilon^2)^2}\left[\epsilon K_1^{[1]}(\infty) \right.
&-&
2\epsilon(1-\epsilon^2/2)
K_{-1}^{[1]}(\infty)\left\langle V^2\right\rangle_\infty+\epsilon^2K_{-2}^{[1]}(\infty)
\left\langle V^3\right\rangle_\infty
\nonumber\\
& & \left.
-\sum_{r\geq 4}\epsilon^{r-1}{2\over r!}\left(r+\epsilon^2(r-3)\right)K^{[1]}_{-r+1}
(\infty)
\left\langle V^r\right\rangle_\infty\right].
\label{heatstation}
\end{eqnarray}
Taking into account (\ref{v2}), Eq.(\ref{heatstation}) implies at order $\epsilon$
\begin{equation}
\left\langle w^3\right\rangle_\infty^{[1]}\left\langle w\right\rangle_\infty^{[2]}=
\left\langle w^3\right\rangle_\infty^{[2]}\left\langle w\right\rangle_\infty^{[1]}.
\end{equation}
Assuming Maxwellian distribution for the fluid particles,
it implies $T_{[1]}(\infty)=T_{[2]}(\infty)$ at order $\epsilon$
and in this case we have
\begin{equation}
\phi_{[1]}(v)=\phi_{[2]}(-v).
\label{equaldistr}
\end{equation}
We conjecture that for any equilibrium state
the relation (\ref{equaldistr}) must be satisfied,
which then implies $T_{[1]}(\infty)=T_{[2]}(\infty)$.
Assuming this conjecture (\ref{equaldistr}) to hold, Eq.(\ref{equal1}) yields
first
\begin{eqnarray}
&& K_0^{(l)}(\infty)=0\quad {\rm for}\quad l=0,1\\
{\sl i.e.} \quad && n_{[1]}^{(l)}(\infty)= n_{[2]}^{(l)}(\infty)
\quad {\rm for}\quad l=0,1
\label{n1n2}
\end{eqnarray}
Then, with Eq.(\ref{equal2}), Eq.(\ref{n1n2}) is also valid for $l=3$, thus
\begin{equation}
K_r^{(l)}(\infty)=0\quad {\rm for\ all\ }r{\rm\ even},\quad l=0,1,2
\end{equation}
and with (\ref{101})
\begin{equation}
\left\langle V^{2k+1}\right\rangle_\infty^{(1)}=0 \quad{\rm for\ all\ }k\geq 0.
\end{equation}
Iterating the argument at successive order we arrive at the following
conclusion. Under the assumption (\ref{assumption2}) and the conjecture
(\ref{equaldistr}), any equilibrium state of the finite system must satisfy
$n_{[1]}(\infty)=n_{[2]}(\infty)$,
in other words $p_{[1]}(\infty)=p_{[2]}(\infty)$
and $T_{[1]}(\infty)=T_{[2]}(\infty)$.
Conversely, if $\phi_{[1]}(v)=\phi_{[2]}(-v)$, thus $T_{[1]}=T_{[2]}$,
and $n_{[1]}(\infty)
=n_{[2]}(\infty)$ then the solution of
the stationary equation at order $\epsilon$ is
\begin{eqnarray}
\left\langle V^{2k+1}\right\rangle_\infty&=& 0\nonumber\\
\left\langle V^{2k}\right\rangle_\infty &=& (2k-1)!!\Delta^k,\quad \Delta={K_1(\infty)
\over 2K_{-1}(\infty)}.
\end{eqnarray}
Therefore at order $\epsilon$ the velocity distribution of the piston
is necessarily Maxwellian with temperature $T=M\Delta/k_B$,
{\sl i.e.}
\begin{equation}
\Phi_\epsilon(V) =\sqrt{{1\over 2\pi\Delta}}\exp\left[-{V^2\over 2\Delta}\right].
\end{equation}
In conclusion the adiabatic piston is stricto senso a conductor since the
only equilibrium states are those for which $T_{[1]}(\infty)
=T_{[2]}(\infty)$.
However the important
question to be considered in the next section
concerns the time scale necessary to reach this true equilibrium
state.
\section{Time evolution}\label{timeevol}
In this section we give a qualitative (non rigorous) discussion of the time
evolution for the infinite and the finite systems.
\subsection{Infinite systems}
As discussed in Sec.\ref{micro}, for infinite systems and $\epsilon\ll 1$
the assumptions $n_{[i]}(t)=n_{[i]}$, $\phi_{[i]}(v,t)
=\phi_{T_{[i]}}(v)$, $i=1,2$ are justified. In this case the functions
$K_r$ are constant, and we are led to solve the equation (\ref{serie})
\begin{equation}
{d\over dt}{\cal V}=\epsilon\left[{\cal C}+{\cal AV}\right],
\quad {\cal V}(t=0)=0,
\end{equation}
where
\begin{eqnarray}
{\cal V}&=&\left\{\left\langle V^s\right\rangle_t;\ s\geq 1\right\}\nonumber\\
{\cal C}_s&=&{1\over \epsilon}A_{s,0}=2^s\epsilon^{s-1}M^{-{s+1\over 2}}\left[
n_{[1]}\left\langle w^{s+1}\right\rangle^{[1]}+n_{[2]}\left\langle w^{s+1}\right\rangle^{[2]}\right]\nonumber\\
{\cal A}_{s,r}&=&{\epsilon^{|s-r-1|}\over (1+\epsilon^2)^s}{1\over \epsilon}P_{s,r}(\epsilon^2)
K_{s-r-1}
\end{eqnarray}
{\sl i.e.} with $t'=\epsilon t$
\begin{equation}
{d\over dt'}{\cal V}={\cal C}+{\cal AV}.
\end{equation}
In the limit $\epsilon\to 0$, with the condition that
\begin{equation}
\lim_{\epsilon\to 0}{1\over \epsilon}K_0=\lim_{\epsilon\to 0}{1\over \epsilon}\left[p_{[1]}(\epsilon)-
p_{[2]}(\epsilon)\right]
\end{equation}
exists and is finite, the matrix ${\cal A}_{s,r}$ is non-zero
only for $r=s-2,s-1,s$, and its eigenvalues $\alpha_n$ are all negative
given by
\begin{equation}
\alpha_s=-s4K_{-1}=-s{\lambda\over \epsilon M},\quad s\geq 1
\end{equation}
where the friction coefficient $\lambda=\lambda(V=0)$, given by (\ref{friction}), is
of order $\epsilon$.
\noindent
We thus obtain at lowest order in $\epsilon$
\begin{eqnarray}
\left\langle V\right\rangle^{(0)}_t &=& V_0\left(1-{\rm e}^{-{\lambda\over M}t}\right)\nonumber\\
\left\langle V^2\right\rangle^{(0)}_t &=& V_0^2\left(1-{\rm e}^{-{\lambda\over M}t}\right)^2
+\Delta\left(1-{\rm e}^{-2{\lambda\over M}t}\right)
\end{eqnarray}
with $V_0$ and $\Delta$ given by Eqs.(\ref{91},\ref{v2}) or
(\ref{91},\ref{v2}).
\noindent
At order $\epsilon^2$ we find for
the case $p_{[1]}=p_{[2]}$
\begin{eqnarray}
\left\langle V\right\rangle_t &=& \left\langle V\right\rangle_\infty
\left(1-{\rm e}^{-{\lambda\over M(1+\epsilon^2)}t}\right)^2\label{139}\\
\left\langle V^2\right\rangle_t &=&\left\langle V^2\right\rangle_\infty
\left[1-{\rm e}^{-2{\lambda\over M(1+\epsilon^2)}t}+2\epsilon^2
\left({\rm e}^{-{\lambda\over M(1+\epsilon^2)}t}-{\rm e}^{-2{\lambda\over M(1+\epsilon^2)}t}
\right)\right]
\end{eqnarray}
with $\left\langle V\right\rangle_\infty$ and $\left\langle V^2\right\rangle_\infty$ given by Eqs.(\ref{813},
\ref{814}) with $p_{[1]}=p_{[2]}$.
Similarly for $p_{[1]}\ne p_{[2]}$ we will have another linear combination of
$\exp(-t/\tau_a)$, $\exp(-2t/\tau_a)$ with
\begin{equation}
\tau_a={M+m\over \lambda}.
\label{7.1.1}
\end{equation}
We notice that taking realistic numbers one finds $\tau_a$ of the order
$10^{-1}-10^{-2}$ which means that the stationary state is reached
very rapidly.
\subsection{Finite systems}
\label{7.2}
For finite systems the equations for the time evolution of
$\left\langle V\right\rangle_\infty$ and $\left\langle V^2\right\rangle_\infty$ are given at order
$\epsilon^2$ by Eqs.(\ref{701},\ref{702}) where the densities $n_{[i]}$ and
the coefficients $K_r^{[i]}$ are now functions of time.
Since we do not consider the full many-body problem describing the fluids,
we take care of this unknown time dependence by assuming that the
fluids satisfy the homogeneity condition (\ref{homogene}). Moreover for
$K_r^{[i]}$, $r=-2,-1,1$, we take the expressions (\ref{71a}) obtained
with the Maxwellian distributions. From Eq.(\ref{evolt}) we obtain
at order $\epsilon$
\begin{eqnarray}
{d\over dt}T_{[1]} &=& -{2T_{[1]}\over \left\langle X\right\rangle_t}\left[
\left\langle V\right\rangle_t-\sqrt{m}\sqrt{{8\over \pi k_BT_{[1]}}}\left\langle V^2\right\rangle_t
+{\sqrt{m}\over M}\sqrt{{8k_B T_{[1]}\over \pi}}\right]\label{7.2.1}\\
{d\over dt}T_{[2]} &=& -{2T_{[2]}\over L-\left\langle X\right\rangle_t}\left[
-\left\langle V\right\rangle_t-\sqrt{m}\sqrt{{8\over \pi k_BT_{[2]}}}\left\langle V^2\right\rangle_t
+{\sqrt{m}\over M}\sqrt{{8k_B T_{[2]}\over \pi}}\right]\label{7.2.2}
\end{eqnarray}
At order $\epsilon$ we thus have a system of five O.D.E for the unknown
$\left\langle X\right\rangle_t$, $\left\langle V\right\rangle_t$, $\left\langle V^2\right\rangle_t$, $T_{[1]}$ and $T_{[2]}$.
At this order we conclude that the system will evolve toward the unique
equilibrium state
\begin{eqnarray}
T_{[1]}(\infty)=T_{[2]}(\infty) &=& {N_{[1]}T_{[1]}+N_{[2]}T_{[2]}\over
N_{[1]}+N_{[2]}}\\
\left\langle X\right\rangle_\infty &=& L{N_{[1]}\over N_{[1]}+N_{[2]}}\\
\left\langle V\right\rangle_\infty &=& 0\\
\left\langle V^2\right\rangle_\infty &=& {k_B\over M}{N_{[1]}T_{[1]}+N_{[2]}T_{[2]}\over
N_{[1]}+N_{[2]}}
\end{eqnarray}
which is identical to the one obtained in thermodynamics for the conducting
piston ($\kappa\ne 0$).
Let us then analyze qualitatively how the evolution toward this final
equilibrium state takes place.
Given that at $t=0$, we have $X=X_0$, $V=0$, $T_{[1]}(0)=T_{[1]}$,
$T_{[2]}(0)=T_{[2]}$, and $p_{[1]}(0)\ne p_{[2]}(0)$, then as long
as the velocity $\left\langle V\right\rangle_t$ remains small Eqs.(\ref{7.2.1},\ref{7.2.2})
yield the adiabats
\begin{eqnarray}
T_{[1]}(t)\left\langle X\right\rangle_t^2 &=& T_{[1]}X_0^2\label{7.2.7}\\
T_{[2]}(t)\left(L-\left\langle X\right\rangle_t\right)^2 &=& T_{[2]}\left(L-X_0\right)^2
\label{7.2.8}
\end{eqnarray}
and the time evolution for the piston is
\begin{equation}
{d\over dt}\left\langle V\right\rangle_t={k_B\over M}\left[
N_{[1]}T_{[1]}{X_0^2\over \left\langle X\right\rangle_t^3}
-N_{[2]}T_{[2]}{\left(L-X_0\right)^2\over \left(L-\left\langle X\right\rangle_t\right)^3}
\right]
-{\lambda(t)\over M}\left\langle V\right\rangle_t
\label{149}
\end{equation}
with $\lambda(t)$ given by Eq.(\ref{71a}), together with Eqs.(\ref{7.2.7},
\ref{7.2.8}).
The piston evolves thus adiabatically until a time $t_a$, of the order
$\tau_a$ Eq.(\ref{7.1.1}), where the pressures become equal, {\sl i.e.}
\begin{equation}
{N_{[1]}T_{[1]}(t_a)\over \left\langle X\right\rangle_{t_a}}=
{N_{[2]}T_{[2]}(t_a)\over L-\left\langle X\right\rangle_{t_a}}
\end{equation}
which yields
\begin{eqnarray}
\left\langle X\right\rangle_{t_a} &=&{L\over 1+\left[{N_{[2]}T_{[2]}\over N_{[1]}T_{[1]}}
\left({L-X_0\over X_0}\right)^2\right]^{1/3}}\label{151}\\
T_{[1]}(t_a)&=&T_{[1]}\left({X_0\over \left\langle X\right\rangle_{t_a}}\right)^2\\
T_{[2]}(t_a)&=&T_{[2]}\left({L-X_0\over L-\left\langle X\right\rangle_{t_a}}\right)^2\\
p(t_a)&=&k_BN_{[1]}T_{[1]}{X_0^2\over \left\langle X\right\rangle_{t_a}^3}.\label{154}
\end{eqnarray}
In the next stage, as can be seen from the numerical simulations of
Sec.\ref{simul}, the pressures will remain approximatively constant
and equal, {\sl i.e.}
\begin{eqnarray}
p_{[1]}(t) &=& k_B{N_{[1]}T_{[1]}(t)\over \left\langle X\right\rangle_t}
\simeq {\rm cte}\label{7.2.15}\\
p_{[2]}(t) &=& k_B{N_{[2]}T_{[2]}(t)\over L-\left\langle X\right\rangle_t}\simeq {\rm cte}
\label{7.2.16}\end{eqnarray}
and
\begin{equation}
N_{[1]}T_{[1]}(t)\left[L- \left\langle X\right\rangle_t\right]=N_{[2]}T_{[2]}(t)\left\langle X\right\rangle_t
\label{7.2.17}
\end{equation}
From Eqs.(\ref{7.2.15},\ref{7.2.16}) we have
\begin{eqnarray}
{1\over T_{[1]}}{d\over dt}T_{[1]} &=& {\left\langle V\right\rangle_t\over \left\langle X\right\rangle_t}\\
{1\over T_{[2]}}{d\over dt}T_{[2]} &=& {\left\langle V\right\rangle_t\over L-\left\langle X\right\rangle_t}
\end{eqnarray}
and using Eqs.(\ref{7.2.1},\ref{7.2.2}) we conclude that
\begin{eqnarray}
M\left\langle V^2\right\rangle_t &=& k_B \sqrt{T_{[1]}(t)T_{[2]}}(t)\label{7.2.20}\\
\left\langle V\right\rangle_t &=& {2\over 3} {\sqrt{m}\over M}\sqrt{{8k_B\over \pi}}\left[
\sqrt{T_{[2]}(t)}-\sqrt{T_{[1]}(t)} \right].
\label{7.2.21}
\end{eqnarray}
In other words the temperatures $T_{[1]}(t)$ and $T_{[2]}(t)$
evolves, but at all time $t$ we have
\begin{eqnarray}
\left\langle V^2\right\rangle_t &=&\left\langle V^2\right\rangle_\infty\\
\left\langle V\right\rangle_t &=&{16\over 3\pi}\left\langle V\right\rangle_\infty
\end{eqnarray}
where $\left\langle V\right\rangle_\infty$ and $\left\langle V^2\right\rangle_\infty$ are the stationary
values for the infinite system with temperatures $T_{[i]}=T_{[i]}(t)$,
$i=1,2$.
\noindent
Introducing the expressions (\ref{7.2.20}) and (\ref{7.2.21}) in
(\ref{7.2.1}), (\ref{7.2.2}) yields, with $p_{[1]}=p_{[2]}=p$
\begin{eqnarray}
{d\over dt}T_{[1]}&=&-{2\over 3}{p\over N_{[1]}k_B}{\sqrt{m}\over M}
\sqrt{{8k_B\over \pi}}\left[\sqrt{T_{[1]}}-\sqrt{T_{[2]}}\right]
\label{7.2.22}\\
{d\over dt}T_{[2]}&=&{2\over 3}{p\over N_{[2]}k_B}{\sqrt{m}\over M}
\sqrt{{8k_B\over \pi}}\left[\sqrt{T_{[1]}}-\sqrt{T_{[2]}}\right]
\end{eqnarray}
and thus
\begin{equation}
{d\over dt}(T_{[2]}-T_{[1]})=-{2\over 3}\left({1\over N_{[1]}k_B}+
{1\over N_{[2]}k_B}\right)
{\sqrt{m}\over M}
\sqrt{{8k_B\over \pi}}
{n_{[1]}k_B T_{[1]}\over \sqrt{T_{[1]}}+\sqrt{T_{[2]}}}
\left(T_{[2]}-T_{[1]}\right).
\end{equation}
Let us note that for $p_{[1]}=p_{[2]}$, it follows from
Eqs.(\ref{conductivity},\ref{friction}) that
\begin{eqnarray}
\kappa &=& {\sqrt{m}\over M}
\sqrt{{8k_B\over \pi}}
{n_{[1]}k_B T_{[1]}\over \sqrt{T_{[1]}}+\sqrt{T_{[2]}}}\\
\lambda_{[1]}+\lambda_{[2]}&=&\sqrt{m}\sqrt{{8k_B\over \pi}}
n_{[1]}T_{[1]}{\sqrt{T_{[1]}}+\sqrt{T_{[2]}}\over \sqrt{T_{[1]}T_{[2]}}}\\
{\sl i.e.}\quad \kappa &=& {\lambda\over M}k_B(\theta+\theta^{-1})^2,\quad
\theta=\left({T_{[1]}\over T_{[2]}}\right)^{1/4}.
\end{eqnarray}
We thus have
\begin{equation}
{d\over dt}(T_{[2]}-T_{[1]})=-{2\over 3}
\left({1\over N_{[1]}k_B}+
{1\over N_{[2]}k_B}\right)\kappa\left(T_{[2]}-T_{[1]}\right)
\label{170}
\end{equation}
from which follows that the system will reach equilibrium with a
relaxation time $\tau_{\rm e}$ of the order
\begin{equation}
\tau_{\rm e}\simeq \kappa^{-1}\simeq {1\over k_B}{M\over \lambda}
\simeq {1\over k_B} \tau_a.
\end{equation}
Taking realistic numbers we obtain a relaxation time which is several
time the age of the universe.
The coefficient $2\left({1\over N_{[1]}k_B}+{1\over N_{[2]}k_B}\right)$
is clearly the inverse of the specific heat. To understand the
origin of the factor $1/3$ we must remember that the piston is moving.
Using Eqs.(\ref{7.2.21},\ref{7.2.22}) yields
\begin{equation}
P_Q^{[2]\to [1]}={1\over 2}N_{[1]}k_B {dT_{[1]}\over dt}
+p_{[1]}\left\langle V\right\rangle_t=\kappa \left(T_{[2]}-T_{[1]}\right)=-P_Q^{[1]\to [2]}
\end{equation}
as it should.
Let us also note that the equality of pressure (\ref{7.2.17}), together
with the conservation of energy
\begin{equation}
N_{[1]}T_{[1]}(t)+N_{[2]}T_{[2]}(t)+{M\over k_B}\left\langle V^2\right\rangle_t=
N_{[1]}T_{[1]}+N_{[2]}T_{[2]},
\end{equation}
which is exactly satisfied by
Eqs.(\ref{701},\ref{702},\ref{7.2.1},\ref{7.2.2}) if $p_{[1]}=p_{[2]}$,
gives the relation between position and temperatures
(\ref{7.2.15},\ref{7.2.16})
\begin{eqnarray}
LN_{[1]}T_{[1]}(t) &=& \left(N_{[1]}T_{[1]}+N_{[2]}T_{[2]}
-{M\over k_B}\left\langle V^2\right\rangle_t\right)\left\langle X\right\rangle_t\nonumber\\
&\simeq &\left(N_{[1]}T_{[1]}+N_{[2]}T_{[2]}\right)\left\langle X\right\rangle_t\\
LN_{[2]}T_{[2]}(t)
&\simeq &\left(N_{[1]}T_{[1]}+N_{[2]}T_{[2]}\right)\left(L-\left\langle X\right\rangle_t
\right)
\end{eqnarray}
and thus from Eq.(\ref{7.2.21}), or Eq.(\ref{7.2.22}),
\begin{equation}
{d\over dt}\left\langle X\right\rangle_t={2\over 3}{\sqrt{m}\over M}
\sqrt{{8\over \pi}}\sqrt{{k_B \left(N_{[1]}T_{[1]}+N_{[2]}T_{[2]}\right)
\over L}}\left(\sqrt{{L-\left\langle X\right\rangle_t\over N_{[2]}}}-
\sqrt{{\left\langle X\right\rangle_t\over N_{[1]}}}\right).
\label{7.2.32}
\end{equation}
Finally if we compare ${d\over dt}\left\langle V\right\rangle_t$ obtained
from Eq.(\ref{7.2.32}) with Eq.(\ref{701}), we arrive at the conclusion
that
\begin{equation}
p_{[2]}-p_{[1]}={\cal O}\left(\epsilon^2(T_{[2]}-T_{[1]})\right)
\end{equation}
which justifies our starting point for the second stage.
Let us also remark that for some time interval
between the adiabatic evolution Eq.(\ref{149}), and
the approach toward equilibrium with $p_{[1]}\simeq p_{[2]}$ (\ref{170}),
there will be an intermediate stage for which $P_Q^{[1]\to[2]}+P_Q^{[2]\to[1]}$
is not zero. During this intermediate evolution the stochastic motion of the
piston will have to be introduced in the thermodynamical equations not only
with a conductivity coefficient $\kappa\ne 0$, but also by means of an
internal energy $U_p$ and and entropy $S_{p}$ of the piston with
$U_p=U_p(S_p)$.
\section{Numerical simulations}\label{simul}
In order to verify the assumptions on which our previous analysis is based,
we made numerical simulations for the finite as well as for
the infinite system in one dimension taking $k_B=1$. The initial state of the
fluids particles of mass $m$ is
given by Maxwellian distributions of velocities according to Eq.(\ref{1}).
Then the particles and the piston (a particle of mass $M$ with initial
coordinate $(X_0,V_0)$) interact only through elastic collisions.
For the infinite system, we simulate the openness of the system
with sources of in-going particles very far from the piston position.
We compute the average time dependent position $\langle X\rangle_t$
of the piston on $10^3-10^4$ different samples.
An example of the time evolution of the piston average position
is shown on Fig. \ref{fig2} for the following parameter
$m=1$, $T_{[1]}=1$, $T_{[2]}=10$, $n_{[1]}=1$, $n_{[2]}=1/10$, {\sl i.e.}
equal pressure on both sides of the piston,
and $M=2,5,10,20,50,100$ (from top to bottom).
The initial position and velocity of the piston are
$X_0=V_0=0$. As expected from Eq.(\ref{139})
we observe that the average position quickly behaves
as $\langle X\rangle_t=\left\langle V\right\rangle_\infty t$ ($t\gtrsim 10^2$).
The relaxation time $\tau_a$ necessary
to reach this stationary behavior is too short to be represented on
Fig. \ref{fig2}, and is presented on Fig. \ref{fig2b} for $M=5,10,50,100$.
Remark that $\tau_a$ depends on the ratio $m/M$.
\begin{figure}
\epsfxsize=11truecm
\hspace{2.75truecm}
\epsfbox{fig2.eps}
\caption{
The average position of the piston as a function of time
for the infinite system. The parameters are
$m=1$, $T_{[1]}=1$, $T_{[2]}=10$, $n_{[1]}=1$, $n_{[2]}=1/10$ and
$M=2,5,10,20,50,100$ (from top to bottom). The dotted box is enlarged
on Fig. \ref{fig2b}.
\label{fig2}}
\end{figure}
\begin{figure}
\epsfxsize=11truecm
\hspace{2.75truecm}
\epsfbox{fig2b.eps}
\caption{
The average position of the piston as a function of time
for the infinite system. Magnification of the dotted box on Fig. \ref{fig2}.
The parameters are as in Fig. \ref{fig2} with
$M=5,10,50,100$.
\label{fig2b}}
\end{figure}
\noindent
Even if the pressures
are equal on both sides of the piston,
{\sl i.e.} $n_{[1]}T_{[1]}=n_{[2]}T_{[2]}$, the stationary velocity
$\left\langle V\right\rangle_\infty$ is non zero
and is presented as a function of $M$ on
Fig. \ref{fig3} for the above parameters, as well as
for $T_{[2]}=20$ and $n_{[2]}=1/20$.
The lines on Fig. \ref{fig3} show the average stationary velocity computed
in Sec.\ref{6.2} up to the fourth order in $\epsilon=\sqrt{m/M}$
\begin{equation}
\left\langle V\right\rangle_\infty=\epsilon\left\langle V\right\rangle_\infty^{(1)}+\epsilon^3\left\langle V\right\rangle_\infty^{(3)}
+{\cal O}(\epsilon^5)
\end{equation}
with (see Eq.(\ref{102}))
\begin{equation}
\left\langle V\right\rangle_\infty^{(1)}
={\sqrt{2\pi}\over 4}{1\over \sqrt{M}}\left(\sqrt{T_{[2]}}
-\sqrt{T_{[1]}}\right)
\end{equation}
and
\begin{equation}
\left\langle V\right\rangle_\infty^{(3)}
={\sqrt{2\pi}\over 96}{1\over \sqrt{M}}\left[(16-3\pi){\left(\sqrt{T_{[2]}}
-\sqrt{T_{[1]}}\right))^3\over
\sqrt{T_{[1]}T_{[2]}}}-6\left(\sqrt{T_{[2]}}-\sqrt{T_{[1]}}\right)\right].
\end{equation}
One sees that the values obtained at the fourth order agree very well
with the numerical results.
\begin{figure}
\epsfxsize=11truecm
\hspace{2.75truecm}
\epsfbox{fig3.eps}
\caption{
The stationary average velocity reached by the piston for the infinite system
as a function of its mass $M$.
The parameters of the simulation are $m=1$, $T_{[1]}=1$, $n_{[1]}=1$ while
$T_{[2]}=10$, $n_{[2]}=1/10$ (circles) and $T_{[2]}=20$, $n_{[2]}=1/20$
(squares), respectively. The solid, resp. dashed,
line shows the theoretical value obtained in the text up to fourth
order in $\epsilon$.
\label{fig3}}
\end{figure}
Concerning the finite system, we present here a qualitative study of the
piston properties as a function of the parameter $\epsilon$.
We consider a finite system of size $L=6\times 10^5$
with $N_{[1]}=N_{[2]}=10^5$ and with
$X_0=10^5$ (thus we have $n_{[1]}=1$, $n_{[2]}=1/5$).
The initial temperatures
on the left and on the right of the piston are $T_{[1]}=1$ and $T_{[2]}=10$.
The initial pressure difference is
thus $p_{[1]}-p_{[2]}=k_B(n_{[1]}T_{[1]}-n_{[2]}T_{[2]})<0$,
pushing the piston toward the left side.
The time scale of our simulations has to be compared with the typical time
it takes for a given particle to cross the system which is of the order
${\cal O}(10^5)$.
We plot on Fig. \ref{fig4} the
time behavior of the piston's position for different values of
$\epsilon=2^{-1/2},10^{-n/2}$ ($n=1,2,3,4$).
\begin{figure}
\epsfxsize=11truecm
\hspace{2.75truecm}
\epsfbox{fig4.eps}
\caption{
The piston's position as a function of time.
The parameters of the simulation are $m=1$, $T_{[1]}=1$, $n_{[1]}=1$ while
$T_{[2]}=10$, $n_{[2]}=1/5$. $X(t)$ is shown for
for $M=2,10,10^2,10^3,10^4$ (from top to bottom).
The dashed line stands for the equilibrium solution
($X=3\times 10^5$) while the dotted lines show the
thermodynamical prediction ($X=84260.5$) for the adiabatic piston
($\kappa=0$).
\label{fig4}}
\end{figure}
\noindent
On Fig. \ref{fig5} and Fig. \ref{fig6}, we plotted
respectively the evolution of the temperatures $T_{[i]}(t)$,
$(i=1,2)$ and of the pressures
for the same parameters.
Notice that the temperatures are estimated through the second moment
of the fluid particles velocity for all the particles present on the same
side of the piston. It is equivalent to the temperatures on the
face of the piston only if we suppose that the system is homogeneous.
The dashed lines show the equilibrium solution predicted in
Sec.\ref{equilfinite} where
$n_{[1]}(\infty)=n_{[2]}(\infty)=(N_{[1]}+N_{[2]})/L$,
$X_{\infty}=LN_{[1]}/(N_{[1]}+N{[2]})$, $T_{[1]}(\infty)=T_{[2]}(\infty)=
(T_{[1]}+T_{[2]})/2$, $p_{[1]}(\infty)=p_{[2]}(\infty)=k_B (N_{[1]}+N_{[2]})
(T_{[1]}+T_{[2]})/(2L)$.
On the other hand, the dotted lines show the equilibrium values obtained
from a numerical integration of
the phenomenological thermodynamical equations (\ref{eqv},\ref{42},\ref{43})
with $\kappa=0$ for the primitive model of an adiabatic piston and
where we used $\lambda_{[i]}(V)\simeq
\lambda_{[i]}(V=0)=n_{[i]}(t)\sqrt{8k_B T_{[i]} m/\pi}$. In this case the system
evolves toward a metastable state with $p_{[1]}(\infty)=p_{[2]}(\infty)$ but
$T_{[1]}(\infty)\ne T_{[2]}(\infty)$. One should note that on the scale
of Figs. \ref{fig4}, \ref{fig5} and \ref{fig6}, one can hardly distinguish
the results obtained by numerical integration ($X\simeq 84260.5$,
$T_{[1]}\simeq 1.545$, $T_{[2]}\simeq 9.455$ and $p_{[1]}=p_{[2]}\simeq 1.83$)
from those obtained
by heuristic arguments Eqs.(\ref{151}-\ref{154}),
$X\simeq 82200$,
$T{[1]}\simeq 1.48$, $T_{[2]}\simeq 9.32$ and $p_{[1]}=p_{[2]}\simeq 1.80$.
\begin{figure}
\epsfxsize=11truecm
\hspace{2.75truecm}
\epsfbox{fig5.eps}
\caption{
The temperature of the fluids on both sides of the piston
as a function of time.
The parameters of the simulation are $m=1$, $T_{[1]}=1$, $n_{[1]}=1$ while
$T_{[2]}=10$, $n_{[2]}=1/5$. $T_{[1]}(t)$, resp. $T_{[2]}(t)$, is shown for
for $M=2,10,10^2,10^3,10^4$ (from middle to bottom, resp. to top).
The dashed line stands for the equilibrium solution
($T_{[1]}(\infty)=T_{[2]}(\infty)=(T_{[1]}+T_{[2]})/2=5.5$) while the
dotted lines show the
thermodynamical prediction ($T_{[1]}(\infty)\sim 1.545 $,
$T_{[2]}(\infty)\sim 9.455$) for $\kappa=0$.
\label{fig5}}
\end{figure}
\begin{figure}
\epsfxsize=11truecm
\hspace{2.75truecm}
\epsfbox{fig6.eps}
\caption{
The pressures of the fluids on both sides of the piston
as a function of time.
The parameters of the simulation are $m=1$, $T_{[1]}=1$, $n_{[1]}=1$,
$p_{[1]}=1$ while
$T_{[2]}=10$, $n_{[2]}=1/5$, $p_{[2]}=2$.
$p_{[1]}(t)$, resp. $p_{[2]}(t)$, is shown for
for $M=10,10^2,10^3,10^4$.
The dashed line stands for the equilibrium solution which is equal to
the thermodynamical prediction
($p_{[1]}(\infty)=p_{[2]}(\infty)=N_{[1]}(T_{[1]}+T_{[2]})/L=11/6$).
\label{fig6}}
\end{figure}
\noindent
Let us first comment on the Fig. \ref{fig6} where one sees that the
pressures quickly become constant and equal to their final values.
The time scale $t_a$ needed to equalize the pressures is roughly independent
on the mass ratio $m/M$. This behavior has been used
in Sec.\ref{7.2} to give a qualitative discussion of the evolution.
Consider now the Fig. \ref{fig5} where one sees that the temperatures reach
quickly an intermediate stage characterized by the equality of the pressures
(time scale $t_a$) then evolve slowly toward their equilibrium values on
a time scale $\tau_{\rm e}$. The time scale $\tau_{\rm e}$
depends strongly on the mass ratio $\epsilon^2=m/M$ and for
$\epsilon\ll 1$ we have $\tau_{\rm e}\gg t_a$. One remarks that, for times
covered in our simulations and
for $\epsilon^2\simeq 10^{-4}$ the system behaves as if the
piston was ``truly'' adiabatic ($\kappa=0$).
We plotted on Fig. \ref{fig7} and Fig. \ref{fig8} the velocity distribution
for the particles of both the fluids. It is averaged on all
particles located on the same side of the piston. The initial conditions are
$T_{[1]}=1$, $T_{[2]}=10$, $p_{[1]}=p_{[2]}=1$,
$X_0=V_0=0$, $m=1$ and $M=10$.
The dotted lines show the Maxwellian
initial distribution $\phi_{T_{[i]}}(v)$.
We present these distributions for three different times
$t=t_1,t_2,t_3$.
\begin{figure}
\epsfxsize=11truecm
\hspace{2.75truecm}
\epsfbox{fig7.eps}
\caption{
The particles velocity distribution $\phi_{[1]}(v,t)$
of the fluid $[1]$ at different time for the finite system.
The initial conditions are
$T_{[1]}=1$, $T_{[2]}=10$, $p_{[1]}=p_{[2]}=1$, $X_0=V_0=0$, $m=1$ and $M=10$.
The dotted line show the initial distribution $\phi_{T_{[1]}}(v)$ while the
dashed line show the expected equilibrium distribution
$\phi_{T_{[1]}(\infty)}(v)$ with $T_{[1]}(\infty)=(T_{[1]}+T_{[2]})/2$.
The squares show the velocity distribution for $t_1=2\times 10^5$ where the
temperature of the fluid is $T_{[1]}(t_1)\simeq 1.6$.
The solid line is for $t_2=2.5\times 10^6$ with $T_{[1]}(t_2)\simeq 3.7$ while
the circles are for $t_3=7.5\times 10^7$ where
$T_{[1]}(t_3)\simeq T_{[1]}(\infty)$.
\label{fig7}}
\end{figure}
\begin{figure}
\epsfxsize=11truecm
\hspace{2.75truecm}
\epsfbox{fig8.eps}
\caption{
The particles velocity distribution $\phi_{[2]}(v,t)$
of the fluid $[2]$ at different time for the finite system.
The initial conditions are
$T_{[1]}=1$, $T_{[2]}=10$, $p_{[1]}=p_{[2]}=1$, $X_0=V_0=0$, $m=1$ and $M=10$.
The dotted line show the initial distribution $\phi_{T_{[2]}}(v)$ while the
dashed line show the expected equilibrium distribution
$\phi_{T_{[2]}(\infty)}(v)$ with $T_{[2]}(\infty)=(T_{[1]}+T_{[2]})/2$.
The squares show the velocity distribution for $t_1=2\times 10^5$ where the
temperature of the fluid is $T_{[2]}(t_1)\simeq 9.4$.
The solid line is for $t_2=2.5\times 10^6$ with $T_{[2]}(t_2)\simeq 7.3$ while
the circles are for $t_3=7.5\times 10^7$
where $T_{[2]}(t_3)\simeq T_{[2]}(\infty)$.
\label{fig8}}
\end{figure}
\begin{figure}
\epsfxsize=11truecm
\hspace{2.75truecm}
\epsfbox{fig9.eps}
\caption{
The time evolution of the
temperatures $T_{[1]}(t)$ (bottom) and $T_{[2]}(t)$ (up) for the same
initial conditions as for Figs. \ref{fig7} and \ref{fig8}.
The successive vertical dotted
lines show the times $t_1,t_2$ and $t_3$ at which
the velocity distributions plotted on
Figs. \ref{fig7} and \ref{fig8} were determined.
\label{fig9}}
\end{figure}
\noindent
For the time $t_1=2\times 10^5$ the system has not reached the relaxation time
$t_a$ and the pressures are not yet constant.
At the time $t_2$ ($t_a<t_2<\tau_{\rm e}$), the system has constant pressures
but has not reached the equilibrium state.
At time $t_3>\tau_{\rm e}$ the system has reached its final
equilibrium state. We present on Fig. \ref{fig9} the time dependence of the
fluid temperatures. Vertical lines stand for times $t_1,t_2,t_3$.
We see that, for times such as $t<\tau_{\rm e}$,
these distributions are not Maxwellian
but develops peaks or wells for small velocities.
Moreover these intermediate time distribution are clearly not
symmetric.
This facts are the reason
why we did not suppose Maxwellian distributions for the fluid particles in our
previous analytical study. After a sufficiently long time of the
order of the equilibrium relaxation time of the system $\tau_e$,
the distributions are very well fitted with the Maxwellian distribution
and we conjecture that $\phi_{[1]}(v,\infty)=\phi_{[2]}(v,\infty)=
\phi_{T_{[1]}(\infty)}(v)$ with $T_{[1]}(\infty)= (T_{[1]}+T_{[2]})/2=11/2$.
The dashed lines on Fig. \ref{fig7} and Fig. \ref{fig8} show the
conjectured equilibrium distribution.
This last conjecture is valid only for $m<M$ as it has been shown that the
equilibrium distribution in the case of $m=M$ is a simple superposition
of the two initial Maxwellian \cite{sinai}.
\section{Conclusions}\label{conclusion}
Several conclusions can be drawn from this investigation on the stochastic
motion of a piston which is adiabatic when rigidly fixed.
For infinite systems, it has been shown that very rapidly a stationary
state is reached, where the average velocity of the piston is a function of the
pressures and temperatures of the fluids on both sides.
In particular if the pressures are equal but the temperatures different
the final average velocity is non-zero and the piston has a
macroscopic motion toward the high temperature region.
This result is related to the asymmetry in the fluctuations of
the piston, due to the asymmetry of the temperatures on both sides,
which in turn induces both a macroscopic force and a heat
current from one side to the other. In other words in the stationary state
the stochastic piston is a conductor. To obtain a stationary state with zero
average velocity it is necessary to compensate the stochastic force
by a macroscopic difference in the pressures of the fluids.
In this very special state no work is done by the piston on the fluids,
but heat is transferred from one side to the other.
Moreover this state is not a state of mechanical equilibrium since
all the odd moments of the piston velocity, except the first one,
are non zero.
For finite systems it has been shown that the system will always evolve
toward the equilibrium state predicted for a conducting piston where
pressures, temperatures and densities are equal on both
sides of the piston , but the time
needed to reach this final equilibrium state will depend strongly on the
mass ratio $m/M$ and can reach several time the age of the universe for
reasonable numbers. This evolution takes place in two stages.
In the first stage the system reaches rather rapidly (although not as fast as
the time needed to reach the stationary state of the infinite system) a state
of ``metastable equilibrium'' where the pressures are equal
but the temperatures different. This initial evolution corresponds
to an adiabatic evolution (no heat transfer) and the results obtained
from the microscopical theory, from the thermodynamic equations, and from
the numerical simulations are in good agreement.
In the second stage the piston and the temperatures of the fluids
evolve in such a way that the pressures remain approximatively constant
and equal. If $m\ll M$ it appears that at all time the average velocity
of the piston co\"\i ncide with the velocity in the stationary state
of the infinite system with values of temperatures and pressures given by
those of the finite system at that time. This observation can be understood
from the fact that the stationary state of the infinite system
is reached very rapidly. If $m\ll M$ the time needed to reach the
final state is so large that on the numerical simulations the
``metastable equilibrium'' state appear to be stable and the piston behaves
as an adiabatic piston on the time scale considered.
On the other hand for $m\simeq M$ the final equilibrium state is reached rather
rapidly. Moreover it was observed that during the time evolution the
velocity distribution of the fluid particles is not Maxwellian and that
this distribution evolves slowly toward the equilibrium
Maxwellian distribution characterized by the equilibrium temperature.
Finally the conclusions obtained from thermodynamics, from kinetic theory,
and from numerical simulations, are all in good agreements.
There appears to be no violation of the second law and no
paradox involved.
\acknowledgements
One of the authors (Ch. G.) is very much indebted to J. L. Lebowitz and
E. Lieb for stimulating and continuous e-mail discussions and remarks,
which have been used for the writing of the Introduction.
|
\section{Introduction}
Optical Gravitational Lensing Experiment (OGLE) is a long term
project focusing on detection and monitoring of microlensing
events. During second phase of the experiment data are collected with
the dedicated 1.3 m Warsaw telescope at the Las Campanas Observatory, Chile
(Udalski, Kubiak and Szyma\'nski 1997). Due to the intrinsically small
probability of the microlensing phenomenon, the main targets are the densest
stellar fields, namely the Galactic Bulge and Magellanic Clouds
(e.g. Paczy\'nski 1996). Some telescope time, however, was also allocated
to observations of other objects in which gravitational lensing is present,
e.g. distant quasars lensed by foreground galaxies. Given good seeing
at the Las Campanas Observatory and the amount of available telescope time,
the OGLE-II survey can provide good time coverage of the variability in
a few selected multiply imaged lenses. Currently two such systems are monitored
by OGLE: QSO 2237+0305 and HE 1104-1805.
Objects of this kind are particularly interesting for cosmology
through determination of time delays. QSO 2237+0305
is a unique system and consists of a high-redshift quasar
at $z=1.695$ quadruply lensed by a relatively nearby galaxy at
$z=0.039$ (Huchra et al. 1985). Because light rays of the four images
pass through the bulge of the foreground galaxy, the optical depth
to lensing on individual stars could be as large as 0.4--0.9 depending
on the image (Webster et al. 1991, Schneider et al. 1988),
making microlensing very likely in Huchra's lens.
Moreover the high degree of symmetry, short distance to the
lensing galaxy, and small image separations produce predicted time delays
of about 1 day or less (Schneider et al. 1988). Thus, intrinsic variability
can be easily distinguished from microlensing effects. Microlensing has already
been discovered in Huchra's lens by Irwin et al. (1989) and confirmed later.
Nevertheless, despite broad interest in this object, so far the best
light curve consists of only 59 measurements in the R band spread over
a period of about 8 years and represents the combined efforts of many observers
(\char31 stensen et al. 1996, Corrigan et al. 1991).
The traditional approach to photometry of this object, i.e. with modeling
of the underlying galaxy light or iterative subtraction of the PSF components
at the positions of the lensed images, requires exceptional seeing and
spatial sampling generally not available with the 1.3 m telescope.
Standard deconvolution algorithms also require superb data and their
uncertainties are not easy to understand. In this paper we present a
qualitatively different approach. Using the optimal image subtraction algorithm
described in detail by Alard and Lupton (1998), and generalized by
Alard (1999) for the case of spatially variable kernel, it is now possible
to match PSFs of two images without noise amplifying division in Fourier space.
Photometry on difference images is free of many complications associated
with other methods.
In Section~2 we describe observations. Section~3 contains a description
of the method with the details of this particular adaptation
and photometry on subtracted images. We summarize the results in Section~4
and future plans in Section~5.
\section{Observations}
All observations presented in this paper were made with the 1.3 m Warsaw
telescope at the Las Campanas Observatory, Chile, which is operated by
the Carnegie Institution of Washington. The ``first generation'' camera
uses a SITe $2048 \times 2048$ CCD detector with $24 \mu m$ pixels resulting
in 0.417 arcsec/pixel scale. Images of QSO 2237+0305 are taken in normal mode
(still frame) at ``medium'' readout speed with the gain 7.1 $e^-$/ADU
and readout noise of 6.3 $e^-$. For the details of the instrumentation setup
we refer to Udalski, Kubiak and Szyma\'nski (1997).
To assure uniformity of the data sets, all observing sequences within
the OGLE project are programmed and can be executed in batch mode. Good
seeing is essential for ground based measurements of Huchra's lens,
independent of the photometric method. Therefore, observations are
restricted to nights with seeing better than about 1.4 arcsec, preferably
with low sky background. Median FWHM of the seeing disk was 1.3 arcsec in
the data presented here. The observing season for QSO 2237+0305 at LCO lasts
from late April to mid December. In satisfactory weather conditions
on a given night two 4 minute exposures in the standard $V$ photometric band
are taken, shifted by a few arc seconds with respect to each other.
This can be done typically once or twice per week without a significant
slow down of the primary program. In 1997, we obtained 80 points during
the observing season. In 1998, QSO 2237+0305 was temporarily removed from
the observing program and there are only 19 points during that period.
However, starting from 1999 we expect to get time coverage comparable
to the 1997 observing season.
The OGLE-II data pipeline automatically detects newly collected frames
and performs bias and flatfield corrections (Udalski, Kubiak and Szyma\'nski
1997). At this point initially reduced frames may be passed to photometric
reductions, provided that there is automated software for that purpose.
We describe such software in Section~3 and 5 along with the planned alert
system for high magnification events in QSO 2237+0305.
\section{Photometry}
\subsection{Image subtraction and difference photometry}
Systems such as Huchra's lens pose a remarkable level of complication for
ground based photometry with medium seeing. Because of small image
separations and their proximity to the galaxy nucleus, each pixel near
the center of the blend contains a light contribution from each of 4 lensed
components. Fainter, but complicated, light distribution from the underlying
barred spiral only makes things worse, with the most significant contamination
due to the galaxy core. In the OGLE data the positions of all the lensed
components fit within an area of about 25 pixels.
Clearly, a full fit of positions and intensities is out of the question,
even assuming that we knew the shape of the galaxy light distribution.
Top panels of Fig.~1 illustrate the observational situation.
On the other hand, this extreme local crowding occurs in the field with very
low density of stars, which complicates determination of the PSF, especially
in the presence of spatial gradients. Therefore the basic strategy adopted
here is to fix as many parameters as possible and avoid fitting the galaxy
light altogether by means of image subtraction. We subtract a reference image
from each of the exposures and obtain a series of frames with only the variable
part of the flux. This should completely remove the foreground galaxy
and leave a blend of four variable components with known geometry,
each of them PSF shaped.
Some preliminary processing is required before one can subtract two images
of the same stellar field. Both images must be resampled onto the same
coordinate grid and PSFs need to be matched. The first step is extracting
a $250 \times 500$ arcsec field centered on the object. In this relatively
small field the spatial gradients of the PSF are not too large for
a low order fit and at the same time the field contains enough stars to obtain
a reliable subtraction. Stars are detected at maxima of the cross-correlation
function with the approximate Gaussian model of the PSF. Cross-correlation
image is done by convolving with the lowered Gaussian filter.
For the coordinate transformation we use a first order fit to coordinates
of about 12 stars found in both a given frame and the reference frame.
A simple algorithm for detecting and removing cosmic rays is run before
resampling of the processed image to the coordinate system of the reference
image. We use a bicubic spline interpolator to perform the resampling.
At this stage a resampled frame and the reference frame are fed to the main
program which calculates spatially variable convolution kernel between
both frames. This code utilizes a newly developed method by Alard and Lupton
(1998). Further development of the method for spatially variable kernels
(Alard, 1999) is central to the application described in this paper
because of the reasons outlined above.
Briefly, if one decomposes the convolution kernel into N basis functions
with constant shape, the problem reduces to a linear least
squares fit for the amplitudes of all kernel components. Convolutions
of the reference image with each of the N kernel components become
basis vectors and can be used to fit any other image of the same
stellar field provided it has been resampled to the coordinate grid
of the reference image. The original method assumes that there are no spatial
gradients of the PSF. Remarkably, this algorithm works best in crowded fields,
where the majority of pixels contain information about the PSF. It can even
treat some weak PSF gradients if we can afford subdividing the image into
smaller pieces.
The implementation gets complicated when the above assumption of a constant
kernel breaks down. In this case the number of coefficients needed in
order to fit spatial variation of the n-th order is $(n+1)\times(n+2)/2$
larger than in the previous problem and direct extension of the algorithm
induces unrealistic computing requirements. However, by assuming that
spatial variations of the kernel are negligible on the scale of the
kernel size, one can reduce the most time consuming calculation, that
of elements of the least squares matrix, to shuffling elements of the
corresponding (much smaller) matrix from the previous problem with constant
convolution kernel.
We found that the constant part of the kernel solution was well described
by three Gaussians with sigmas 0.7, 1.2 and 1.8 pix modified
by polynomials of orders 4, 3 and 2, respectively. We allow only the first
order spatial variation of the kernel since there are only 9 stars
that can be used in the fit. In the output we have a subtracted image with
the seeing of the currently processed frame and intensity scale corresponding
to the reference frame, plus the best fit convolution kernel. In all difference
images the residuals of constant stars were consistent with the photon noise of
a single frame. The mean of about 20 frames with the best seeing, low sky
background and resampled onto the same coordinate grid is the correct choice
of the reference image and allows approaching the photon noise limit.
We used the 18 best frames to obtain a good reference image, practically
noiseless compared to a single exposure. In difference images the galaxy
is completely removed outside the region dominated by the lens with residuals
symmetric about zero and consistent with the photon noise. We believe that
this is also the case inside this region. Fig.~1 shows central parts of the
reference image and typical test image, as well as the corresponding
difference image and the best fit PSF matching kernel. Sample difference
images from various epochs are shown in Fig.~2. Variability of all quasar
components is obvious.
The difference image contains only the variable part of the flux and can be
used for high precision relative photometry. After the galaxy is removed
remaining light can be modeled with 4 PSFs. The geometry of the lensed quasar
images is known with a very high accuracy from HST data. We adopted positions
relative to component ``A'' from Crane et al. (1991). In numerous difference
frames, ``A'' was the only quasar image which significantly changed its
brightness with respect to the reference frame. Therefore the PSF component
at the position of ``A'', corresponding to the difference flux, was practically
free of contamination by the remaining components of the blend.
This simplified calculation of its centroid in a stack of 40 such frames
and thus obtaining positions of all lensed quasar images.
The last step is a linear fit to the amplitudes of the
four lensed components in the difference image. The area dominated by variable
light and therefore useful for this purpose consists of pixels with centers
not further than 2.4 pix from any of the quasar images.
We modeled the first order spatial variation of the PSF in the reference
image using the code written for the DENIS survey (Alard 1999, in preparation).
For any other frame (and difference image) the PSF is calculated simply
by convolving the reference PSF with the best fit PSF matching kernel at the
position of the measured object.
Errors of the individual photometric points were estimated by propagating
the photon noise through the linear least squares fit and adding in quadrature
a correction for uncertain flatfield at the level of 1\% of the mean amplitude
of ``A'' component. This simple noise model gives error bars consistent with
the scatter of the individual measurements.
\subsection{Zero point calibration}
The procedure described in the previous paragraph gives only a variable
part of the flux, for example in counts per second. Putting the light curve
on a magnitude scale requires the knowledge of the absolute amplitude
of all components and a reference star in at least one image.
In the presence of the intervening galaxy this requires a reasonably good model
of the underlying light distribution.
We used public HST images from the archive at STScI\footnotemark. The images
were taken on June 23, 1995 with the WFPC2 using F555 filter, centered on
$\lambda = 5407.0 ~ \rm \AA$, the closest available band to $V$. Proposal
ID is 5236. Four of the images had exposure times of 200 seconds and
one of them was exposed for 800 seconds.
\footnotetext{NASA/ESA Hubble Space Telescope is operated by AURA
under NASA contract NAS 5-26555}
The galaxy template was prepared by symmetrization with respect to the
brightest pixel. In the annulus contaminated by quasar images symmetric pairs
of pixels are examined and lower of the two is adopted as the best guess at
the value of both pixels with 0.6$\sigma$ correction for the bias due to
minimization. This simplified version of the symmetrization technique
used by the SDSS survey for deblending star and galaxy images
(Lupton 1999, in preparation) removes effectively quasar
components ``A'' and ``B'', however it fails for ``C'' and ``D'', due to their
very symmetric position with respect to the galaxy nucleus.
Fortunately in WFPC2 images galaxy is smooth on a much larger scale than
the area occupied by any of the quasar images. It is possible to make a local
fit to the galaxy light and subtract the faintest image before symmetrization,
which solves the problem.
The galaxy template must be rotated, resampled to the pixel grid of the OGLE
reference frame, and degraded to the seeing of each OGLE test frame before
the fit. We fitted a model consisting of 4 PSFs and the galaxy template
to every image in our data set. This is much like the conventional approach,
except that in the process we used PSF matching kernels obtained with
the image subtraction software. The scatter of this photometry was about
a factor of two larger than in the image subtraction method. Nevertheless,
the weighted mean of the difference between both light curves (in counts)
is the desired amplitude of a given quasar component in the reference image.
The statistical quality of our zero point is about 2, 3, 8 and 13 \% for
components ``A'', ``B'', ``C'' and ``D'' respectively, therefore it is
much worse than the accuracy of the difference signal with respect to the
reference frame. The data were reduced to the standard $V$ magnitudes based
on observations of standard stars from selected fields of Landolt (1992)
obtained on 11 photometric nights. We get $V=17.466\pm0.018$ and
$18.138\pm0.021$ mag for stars $\alpha$ and $\beta$ of Corrigan et al. (1991),
brighter only by about 0.038 mag compared to the photometry obtained indirectly
by these authors.
\newpage
\section{Results}
Fig.~3 shows the light curve of the QSO 2237+0305. Photometry is given in
Table~1. Machine readable data can be obtained from the OGLE web page
(see the last paragraph for the address).
The fifth order polynomial fits to light variations are also shown to guide
the eye. All components display significant variations, especially between
the two observing seasons, and even more importantly all of them varied
differently. The 18.14 mag reference star measured using exactly the same
method was constant within the errors which confirms that our photometry
is correct (Fig.~3). Component ``C'' brightened by as much as 0.7 mag and
actually almost exchanged in brightness with ``B''. The easiest explanation
to these phenomena is microlensing in the bulge of the macro lensing galaxy.
For the first time we see it happening in a sense that smooth
variation of source amplification is observed, most striking for
component ``A''. Wambsganss, Paczy\'nski and Schneider (1990) demonstrated
huge diversity of possible light variations produced by complicated network
of clustered microcaustics and showed that the sharpest features present
in the light curve are directly related to the size of the source,
the masses of the microlenses and the transverse velocity. They also
stressed that frequently sampled light curves should constrain some of
these unknowns. However in this paper we do not attempt any further
theoretical interpretation of the data.
It must be emphasized that difference photometry with respect to the
reference image is very accurate, however the overall normalization
of the light curve can be off by 15 \% for the faintest component.
On magnitude scale this affects the shape of the light curve too, since for
any test image we have:
$$ {\rm m}_V = const. - 2.5 \times
\log\left({{f_{V,\rm ref} + \Delta f_V} \over norm}\right) ~~~ \rm mag $$
In the above formula the difference flux $\Delta f_V$ is known with high
precision, however the amplitude of a given component in the reference
frame $f_{V,\rm ref}$ is known less accurately. For some applications it may
be safer to go back to the linear scale and obtain zero point free shape
of the light curve. This is accomplished for instance if we express flux
density in milli Janskys ($10^{-29}$ Wm$^{-2}$Hz$^{-1}$) :
$$ f_V = f_{V,\rm ref} + \Delta f_V =
3.67 \times 10^{(-0.4 \cdot {\rm m}_V + 6)} ~~~ \rm mJy ,$$
\noindent
as has been done in Fig.~4.
\section{Discussion and future perspective}
We demonstrated that good quality monitoring of QSO 2237+0305 is possible with
1.3 m telescope. Despite the fact that complications characteristic of both
crowded and sparse fields are present, photometry of this system is well
handled by image subtraction method of Alard \& Lupton (1998) and Alard (1999)
supplemented with PSF fitting of the difference image. Resulting light curves
are significantly better than any other preexisting measurements, even with
much larger instruments. Probably the most important improvement is due to
much better time coverage, especially during 1997 observing season.
Beginning from 1999 we expect to obtain similar light curves during entire
period when this object is accessible from LCO. It is likely that the zero
point will be improved in the future using overlapping observations
from instruments with better seeing and/or better galaxy template.
All data processing described in Section~3 has been integrated
to the level, at which it is possible to run reductions automatically
once a new image is collected and simply wait for the new point to be added
to a postscript plot. Photometric pipeline consists of several stand by
programs for each step of reductions, controlled by a shell script.
It is planned that this software will become a part of the OGLE photometric
pipeline at LCO. This will provide an easy way of checking the state
of Huchra's lens in real time and therefore we should be able to detect
caustic crossings early enough to issue an alert. Caustic crossing provides
in principle the way of resolving spatially the source, which in this case
would place a very tight limit on the size of the quasar
(Wambsganss, Paczy\'nski and Schneider 1990). For this measurement it is
essential to get good coverage of a High Amplification Event, with larger
instruments and better instrumental seeing if possible, therefore
the importance of early detection of HAE cannot be overestimated.
Up-to-date information on Huchra's lens is available from the OGLE web page at
http://www.astrouw.edu.pl/$\sim$ftp/ogle and its USA mirror
http://www.astro.princeton.edu/$\sim$ogle.
\acknowledgments{
We would like to thank Prof. Bohdan Paczy\'nski for
pointing to us the importance of this research, numerous useful discussions
and constant support at all stages of the OGLE project. We also thank
Robert Lupton for his insights on image processing. Comments by Dave Goldberg
and Robert Lupton helped us to improve the manuscript. This work was supported
with the Polish KBN grant 2P03D00814 to A. Udalski and
NSF grant AST--9530478 to B. Paczy\'nski.
}
|
\section{#1}\setcounter{equation}{0}}
\def\nappendix#1{\vskip 1cm\no{\bf Appendix
#1}\def#1} \setcounter{equation}{0}{#1} \setcounter{equation}{0}}
\newfam\dlfam \def\dl{\fam\dlfam\tendl}
\newfam\glfam \def\gl{\fam\glfam\tengl}
\def\number\month/\number\day/\number\year\ \ \ \hourmin {\number\month/\number\day/\number\year\ \ \ \hourmin }
\def\pagestyle{draft}\thispagestyle{draft{\pagestyle{draft}\thispagestyle{draft}
\global\def1}} \global\def\draftcontrol{0{1}} \global\def1}} \global\def\draftcontrol{0{0}
\catcode`\@=12 \def\widetilde} \def\hat{\widehat{\widetilde} \def\hat{\widehat}
\documentstyle[12pt,epsf]{article}
\def{\thesection.\arabic{equation}}{{#1} \setcounter{equation}{0}.\arabic{equation}}}
\setlength{\textwidth}{15.6cm} \setlength{\textheight}{23.12cm}
\hoffset -2cm \topmargin= -0.4cm \raggedbottom
\raggedbottom \renewcommand{\baselinestretch}{1.0}
\newcommand{\be}{\begin{eqnarray}} \newcommand{\en}{\end{eqnarray}\vskip} \newcommand{\hs}{\hspace
0.5 cm} \newcommand{\nonumber} \newcommand{\no}{\noindent}{\nonumber} \newcommand{\no}{\noindent}
\newcommand{\vskip} \newcommand{\hs}{\hspace}{\vskip} \newcommand{\hs}{\hspace}
\newcommand{\'{e}} \newcommand{\ef}{\`{e}}{\'{e}} \newcommand{\ef}{\`{e}}
\newcommand{\partial} \newcommand{\un}{\underline}{\partial} \newcommand{\un}{\underline}
\newcommand{\overline}{\overline}
\newcommand{\NR}{{{\bf R}}
\newcommand{\NA}{{{\bf A}}
\newcommand{\NB}{{{\bf B}}
\newcommand{\NP}{{{\bf P}}
\newcommand{\NC}{{{\bf C}}
\newcommand{\NT}{{{\bf T}}
\newcommand{\NZ}{{{\bf Z}}
\newcommand{\NH}{{{\bf H}}
\newcommand{\NM}{{{\bf M}}
\newcommand{\NN}{{{\bf N}}
\newcommand{\NS}{{{\bf S}}
\newcommand{\NW}{{{\bf W}}
\newcommand{\NV}{{{\bf V}}
\newcommand{{{\bf a}}} \newcommand{\Nb}{{{\bf b}}}{{{\bf a}}} \newcommand{\Nb}{{{\bf b}}}
\newcommand{{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}}{{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}}
\newcommand{{{\bf h}}} \newcommand{\Nr}{{{\bf r}}}{{{\bf h}}} \newcommand{\Nr}{{{\bf r}}}
\newcommand{{{\bf j}}} \newcommand{\Nl}{{{\bf l}}}{{{\bf j}}} \newcommand{\Nl}{{{\bf l}}}
\newcommand{{{\bf x}}} \newcommand{\Ny}{{{\bf y}}}{{{\bf x}}} \newcommand{\Ny}{{{\bf y}}}
\newcommand{{{\bf v}}} \newcommand{\Nw}{{{\bf w}}}{{{\bf v}}} \newcommand{\Nw}{{{\bf w}}}
\newcommand{{{\bf u}}} \newcommand{\Ns}{{{\bf s}}}{{{\bf u}}} \newcommand{\Ns}{{{\bf s}}}
\newcommand{{{\bf t}}} \newcommand{\Nz}{{{\bf z}}}{{{\bf t}}} \newcommand{\Nz}{{{\bf z}}}
\newcommand{{{\bf k}}} \newcommand{\Np}{{{\bf p}}}{{{\bf k}}} \newcommand{\Np}{{{\bf p}}}
\newcommand{\Nq}{{{\bf q}}}
\newcommand{\begin{eqnarray}} \newcommand{\de}{\bar\partial}{\begin{eqnarray}} \newcommand{\de}{\bar\partial}
\newcommand{\partial} \newcommand{\ee}{{\rm e}}{\partial} \newcommand{\ee}{{\rm e}}
\newcommand{{\rm Ker}} \newcommand{\qqq}{\end{eqnarray}}{{\rm Ker}} \newcommand{\qqq}{\end{eqnarray}}
\newcommand{\mbox{\boldmath $\lambda$}}{\mbox{\boldmath $\lambda$}}
\newcommand{\mbox{\boldmath $\alpha$}}{\mbox{\boldmath $\alpha$}}
\newcommand{\mbox{\boldmath $x$}}{\mbox{\boldmath $x$}}
\newcommand{\mbox{\boldmath $\xi$}}{\mbox{\boldmath $\xi$}}
\newcommand{\mbox{\boldmath $k$}} \newcommand{\tr}{\hbox{tr}}{\mbox{\boldmath $k$}} \newcommand{\tr}{\hbox{tr}}
\newcommand{\hbox{ad}} \newcommand{\Lie}{\hbox{Lie}}{\hbox{ad}} \newcommand{\Lie}{\hbox{Lie}}
\newcommand{{\rm w}} \newcommand{\CA}{{\cal A}}{{\rm w}} \newcommand{\CA}{{\cal A}}
\newcommand{{\cal B}} \newcommand{\CC}{{\cal C}}{{\cal B}} \newcommand{\CC}{{\cal C}}
\newcommand{{\cal D}} \newcommand{\CE}{{\cal E}}{{\cal D}} \newcommand{\CE}{{\cal E}}
\newcommand{{\cal F}} \newcommand{\CG}{{\cal G}}{{\cal F}} \newcommand{\CG}{{\cal G}}
\newcommand{{\cal H}} \newcommand{\CI}{{\cal I}}{{\cal H}} \newcommand{\CI}{{\cal I}}
\newcommand{{\cal J}} \newcommand{\CK}{{\cal K}}{{\cal J}} \newcommand{\CK}{{\cal K}}
\newcommand{{\cal L}} \newcommand{\CM}{{\cal M}}{{\cal L}} \newcommand{\CM}{{\cal M}}
\newcommand{{\cal N}} \newcommand{\CO}{{\cal O}}{{\cal N}} \newcommand{\CO}{{\cal O}}
\newcommand{{\cal P}} \newcommand{\CQ}{{\cal Q}}{{\cal P}} \newcommand{\CQ}{{\cal Q}}
\newcommand{{\cal R}} \newcommand{\CS}{{\cal S}}{{\cal R}} \newcommand{\CS}{{\cal S}}
\newcommand{{\cal T}} \newcommand{\CU}{{\cal U}}{{\cal T}} \newcommand{\CU}{{\cal U}}
\newcommand{{\cal V}} \newcommand{\CW}{{\cal W}}{{\cal V}} \newcommand{\CW}{{\cal W}}
\newcommand{{\cal X}} \newcommand{\CY}{{\cal Y}}{{\cal X}} \newcommand{\CY}{{\cal Y}}
\newcommand{{\cal Z}} \newcommand{\s}{\hspace{0.05cm}}{{\cal Z}} \newcommand{\s}{\hspace{0.05cm}}
\newcommand{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}}{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}}
\newcommand{{\rightarrow}} \newcommand{\mt}{{\mapsto}}{{\rightarrow}} \newcommand{\mt}{{\mapsto}}
\newcommand{{_1\over^2}}{{_1\over^2}}
\newcommand{{h\hspace{-0.23cm}^-}}{{h\hspace{-0.23cm}^-}}
\newcommand{{\slash\hs{-0.21cm}\partial}} \pagestyle{plain}{{\slash\hs{-0.21cm}\partial}} \pagestyle{plain}
\begin{document}
\title{\bf{Conformal field theory:\\
a case study}}
\author{ \\Krzysztof Gaw\c{e}dzki\\C.N.R.S., I.H.E.S.,
91440 Bures-sur-Yvette, France\\ }
\date{ }
\maketitle
\vskip 0.2cm
\begin{abstract}
\vskip 0.2cm
\noindent This is a set of introductory lecture notes
devoted to the Wess-Zumino-Witten model of two-dimensional
conformal field theory. We review the construction
of the exact solution of the model from the functional
integral point of view. The boundary version of the theory
is also briefly discussed.
\vskip 0.3cm
\hfill
\end{abstract}
\vskip 1cm
\nsection{Introduction}
Quantum field theory is a structure at the root of our
understanding of physical world from the subnuclear scales
to the astrophysical and the cosmological ones. The concept
of a quantum field is very rich and still poorly
understood although much progress have been achieved
over some 70 years of its history. The main problem
is that, among various formulations of quantum field theory
it is still the original Lagrangian approach which is by far
the most insightful, but it is also the least precise way to
talk about quantum fields. The strong point of the Lagrangian
approach is that it is rooted in the classical theory.
As such, it permits a perturbative analysis of the field
theory in powers of the Planck constant and also captures
some semi-classical non-perturbative effects (solitons,
instantons). On the other hand, however, it masks genuinely
non-perturbative effects. In the quest for a deeper
understanding of quantum field theory
an important role has been played by two dimensional models.
Much of what we have learned about nonperturbative phenomena
in quantum field theory has its origin in such models.
One could cite the Thirring model with its anomalous
dimensions and the fermion-boson equivalence to the
sine-Gordon model, the Schwinger model with the confinement
of electric charge, the non-linear sigma model with the
non-perturbative mass generation, and so on.
\vskip 0.3cm
The two-dimensional models exhibiting conformal invariance
have played a specially important role. On one side,
they are not without direct physical importance, describing,
in their Euclidean versions, the long-distance behavior
of the two-dimensional statistical-mechanical systems,
like the Ising or the Potts models, at the second order phase
transitions. On the other hand, the (quantum) conformal field
theory (CFT) models constitute the essential building blocks
of the classical vacua of string theory, a candidate ``theory
of everything'', including quantum gravity. The two-dimensional
space-time plays simply the role of a string world sheet
parametrizing the string evolution, similarly as the
one-dimensional time axis plays the role of a world line
of point particles. The recent developments seem to indicate
that string theory, or what will eventually
emerge from it, provides the appropriate language to talk
about general quantum fields, whence the central place
of two-dimensional CFT in the quantum field theory edifice.
\vskip 0.3cm
Due to the infinite-dimensional nature of the conformal symmetry
in two space-time dimensions, the two-dimensional models of CFT
lend themselves to a genuinely non-perturbative approach
based on the infinite symmetries and the concept of the operator
product expansion \cite{3}. In the present lectures, we shall
discuss a specific model of two-dimensional CFT, the so called
Wess-Zumino-(Novikov)-Witten model (WZW) \cite{4}\cite{5}\cite{6}.
It is an example of a non-linear sigma model
with the classical fields on the space-time taking values
in a manifold which for the WZW model is taken as a group
manifold of a compact Lie group $G$. We shall root our
treatment in the Lagrangian approach and will work slowly
our way towards a non-perturbative formulation.
This will, hopefully, provide a better understanding of
the emerging structure which, to a large extent, is common
to all CFT models. In fact, the WZW theory is a prototype
of general (rational) CFT models which may be obtained
from the WZW one by different variants of the so called
coset construction. In view of the stringy applications,
where the perturbation expansion is built by considering
two-dimensional conformal theories on surfaces of arbitrary
topology, we shall define and study the WZW model on
a general Riemann surface.
\vskip 0.3cm
A word of warning is due to a more advanced audience.
The purpose of these notes is not to present a complete
up-to-date account of the WZW theory, even less of CFT.
That would largely overpass the scope of a summer-school
lecture notes. As a result, we limit ourselves to
the simplest version of the model leaving completely
aside the ramifications involving models with non-simply
connected groups, orbifolds, etc, as well as applications
to string theory. We profit, however, from this simple example
to introduce on the way some of the main concepts
of two-dimensional CFT.
Much of the material presented is not new,
even old, by the time-scale standard of the subject, with
the possible exception of the last section devoted to
the boundary WZW models. The author still hopes that
the following exposition, which he failed to present
at the 1998 Istanbul summer school, may be useful to
a young reader starting in the field.
\vskip 0.3cm
The notes are organized as follows. In Sect.\,\,2,
we discuss a simple quantum-mechanical version of the WZW
model: the quantum particle on a group manifold.
This simple model, exactly solvable by harmonic analysis
on the group, permits to describe many structures similar
to the ones present in the two-dimensional theory
and to understand better the origin of those.
Sect.\,\,3 is devoted to the definition of the action
functional of the WZW model. The action contains
a topological term, which requires a special treatment.
We discuss separately the case of the surfaces without
and with boundary, in the latter case postponing
the discussion of local boundary conditions
to the last section. In Sect.\,\,4, we introduce
the basic objects of the (Euclidean) quantum WZW theory:
the quantum amplitudes taking values in the
spaces of states of the theory and the correlation
functions. We state the infinite-dimensional symmetry
properties of the theory related to the chiral gauge
transformations and to the conformal transformations.
The symmetries give rise to the action of the two copies
of the current and Virasoro algebras in the Hilbert
space of states of the theory constructed with the
use of the representation theory of those algebras.
We discuss briefly the operator product expansions
which encode the symmetry properties of the correlation
functions. Sect.\,\,5 is devoted to the relation
between the WZW theory and the Schr\"{o}dinger picture
quantum states of the topological three-dimensional
Chern-Simons theory. The relation is established
via the Ward identities expressing the behavior
of the WZW correlation functions, coupled to external
gauge field, under the chiral gauge transformations.
We discuss the structure of the spaces of the Chern-Simons
states, the fusion ring giving rise to the Verlinde
formula for their dimensions and their Hilbert-space
scalar product, as well as the Knizhnik-Zamolodchikov
connection which permits to compare the states
for different complex structures.
In particular, we explain how the knowledge of
the scalar product of the Chern-Simons states permits
to obtain exact expressions for the correlation
functions of the WZW theory.
In Sect.\,\,6 we give a brief account of the coset
construction of a large family of CFT models which may be
solved exactly, given the exact solution of the WZW
model. Finally, Sect.\,\,7 is devoted to the WZW theory
with local boundary conditions. Again, for the sake
of simplicity, we restrict ourselves to a simple family
of the conditions that do not break the infinite-dimensional
symmetries of the theory. We discuss how to define the
action functional of the model in the presence of such
boundary conditions and what are the elementary properties
of the corresponding spaces of states, quantum amplitudes
and correlation functions.
\nsection{Quantum mechanics of a particle
on a group}
\subsection{The geodesic flow on a group}
Non-linear sigma models describe field theories with fields
taking values in manifolds. These lectures will be devoted
to a special type of sigma models, known under the name
of Wess-Zumino(-Novikov)-Witten (or WZW) models. They are
prototypes of conformal field theories in two-dimensional
space-time. As such they play a role in string theory
whose classical solutions are built out of two-dimensional
quantum conformal field theory models by a cohomological
construction. Before we plunge, however, into the details
of the WZW theory, we shall discuss a simpler but largely
parallel model in one dimension, i.e. in the domain of mechanics
rather than of field theory.
\vskip 0.3cm
One-dimensional sigma models describe the geodesic flows
on manifolds $M$ endowed with a Riemannian or a pseudo-Riemannian
metric $\gamma$. The classical action for the trajectory $[0,T]\ni
\,\mapsto\,x(t)\in M$ of such a system is
\begin{eqnarray}} \newcommand{\de}{\bar\partial
S(x)\ =\ {_1\over^2}\int_{_0}^{^T}\hspace{-0.1cm}
\gamma_{\mu\nu}(x)\,
{_{dx^\mu}\over^{dt}}\,{_{dx^\nu}\over^{dt}}\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} dt
\label{act}
\qqq
and the classical solutions $\delta S=0$ correspond to
the geodesic curves in $M$ parametrized by a rescaled length.
If $M=\NR^n$, for example, and $\gamma_{\mu\nu}=m\,\delta_{\mu\nu}$,
we obtain the action of the free, non-relativistic particle
of mass $m$ undergoing linear classical motions
$x(t)=x_0+{p\over m}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} t$.
The action (\ref{act}) is not parametrization-invariant but it
may be viewed as a gauged-fixed version of the action
\begin{eqnarray}} \newcommand{\de}{\bar\partial
S_{_P}(x)\ =\ {_1\over^2}\int_{_0}^{^T}\hspace{-0.1cm}
\gamma_{\mu\nu}(x)\,
{_{dx^\mu}\over^{dt}}\,{_{dx^\nu}\over^{dt}}\,\eta^{-{_1\over^2}}\,dt\ +\
{_1\over^2}\int_{_0}^{^T}\hspace{-0.1cm}\eta^{_1\over^2}\, dt\,,
\nonumber} \newcommand{\no}{\noindent
\qqq
where the reparametrization invariance is restored by
coupling the system {\it \`{a} la} Polyakov to the metric
$\eta(t)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} (dt)^2$ on the word-line of the particle. The
$\eta=1$ gauge reproduces then the action (\ref{act}),
whereas extremizing over $\eta$, one obtains the relativistic
action
\begin{eqnarray}} \newcommand{\de}{\bar\partial
S_r(x)\ =\ \int_{_0}^{^T}\hspace{-0.07cm}
\left(\gamma_{\mu\nu}(x)\,
{_{dx^\mu}\over^{dt}}\,{_{dx^\nu}\over^{dt}}\right)^{{_1\over^2}}dt
\nonumber} \newcommand{\no}{\noindent
\qqq
given by the geodesic length of the trajectory.
\vskip 0.3cm
The exact solvability of the geodesic equations can be
achieved in sufficiently symmetric situations. In particular,
we shall be interested in the case when $M$ is a manifold
of a compact Lie group $G$ and when $\gamma$ is a left-right
invariant metric on $G$ given by ${k\over 2}$ times a positive
bilinear $ad$-invariant form $\tr\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}(XY)$
on the Lie algebra\footnote{we use
the physicists' convention in which the exponential map between
the Lie algebra and the group is $X\mapsto\ee^{iX}$} ${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$.
For matrix algebras, as the algebra $su(N)$ of the hermitian
$n\times n$ traceless matrices, the form is given by the
matrix trace in the defining representation, hence
the notation. The positive constant $k$ will play the role
of a coupling constant.
The action (\ref{act}) may be then rewritten as
\begin{eqnarray}} \newcommand{\de}{\bar\partial
S(g)\ =\ -\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{_k\over^4}\int_{_0}^{^T}\hspace{-0.1cm}
\tr\,\, (g^{-1}{_d\over^{dt}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g)^2\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} dt\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}.
\nonumber} \newcommand{\no}{\noindent
\qqq
The variation of the action under the infinitesimal
change of $g$ vanishing on the boundary is
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\delta S(g)\ =\ {_k\over^2}\int_{_0}^{^T}\hspace{-0.1cm}
\tr\,\,(g^{-1}\delta g)\,\,{_d\over^{dt}}(g^{-1}
{_d\over^{dt}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g))\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} dt\,.
\nonumber} \newcommand{\no}{\noindent
\qqq
Consequently, the classical trajectories are solutions of the
equations
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{_d\over^{dt}}(g^{-1}{_d\over^{dt}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g)\ =\ 0\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}.
\label{clequ}
\qqq
The case $G=\NT^n$, where $\NT^n=\NR^n/\NZ^n$ is the n-dimensional
torus, is the prototype of an integrable system whose
trajectories are periodic or quasiperiodic motions with the angles
evolving linearly in time. The case $G=SO(3)$ corresponds to the
symmetric top whose positions are parametrized by rigid rotations.
The classical trajectories solving Eq.\,\,(\ref{clequ})
have a simple form:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
g(t)=g_\ell\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\ee^{it \lambda/k}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g_r^{-1}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},
\label{alts}
\qqq
where $g_{\ell,r}$ are fixed elements in $G$ and $\lambda$
may be taken in the Cartan subalgebra $\Ntt\subset{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$.
For the later convenience, we have introduced the factor
${1\over k}$ in the exponential.
\vskip 0.3cm
The space ${\cal P}} \newcommand{\CQ}{{\cal Q}$ of classical solutions forms the phase space
of the system. It may be parametrized by the initial data
$(g(0),\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} p(0))$ where the momentum\footnote{we identify ${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$
with its dual using the bilinear form $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\tr\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}(XY)$}
$\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} p(t)={k\over{2i}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{d\over{dt}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g$. \hspace{0.025cm}} \newcommand{\ch}{{\rm ch} As usually,
the phase space ${\cal P}} \newcommand{\CQ}{{\cal Q}$ comes equipped with the symplectic form
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\Omega\ =\ {_1\over^i}\, d\,\tr\,(p\,g^{-1}dg)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},
\qqq
where the right hand side may be calculated at any instance
of time giving a result independent of time. The symplectic
structure on ${\cal P}} \newcommand{\CQ}{{\cal Q}$ allows to associate the vector fields ${\cal X}} \newcommand{\CY}{{\cal Y}_f$
to functions $f$ on ${\cal P}} \newcommand{\CQ}{{\cal Q}$ by the relation $\,-\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} df=\iota_{{\cal X}} \newcommand{\CY}{{\cal Y}_f}
\Omega\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$, \,where $\iota_{X}\alpha$ denotes the contraction
of a vector field ${\cal X}} \newcommand{\CY}{{\cal Y}$ with a differential form $\alpha$.
These are the Hamiltonian vector fields that preserve
the symplectic form:
$\,{\cal L}} \newcommand{\CM}{{\cal M}_{X_f}\Omega=0\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},$ \,where ${\cal L}} \newcommand{\CM}{{\cal M}_{{\cal X}} \newcommand{\CY}{{\cal Y}}$ is the Lie
derivative that acts on differential forms by $\,{\cal L}} \newcommand{\CM}{{\cal M}_{{\cal X}} \newcommand{\CY}{{\cal Y}}\alpha
=\iota_{{\cal X}} \newcommand{\CY}{{\cal Y}}d\alpha+d\iota_{{\cal X}} \newcommand{\CY}{{\cal Y}}\alpha\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$. The Poisson bracket
of functions is defined by: $\{f,g\}={\cal X}} \newcommand{\CY}{{\cal Y}_f(g)$. In particular,
the time evolution is induced by the vector field associated
with the classical Hamiltonian
\begin{eqnarray}} \newcommand{\de}{\bar\partial
h\ =\ {_1\over^k}\,\tr\,\, p^2\ =\ -\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{_k\over^4}\,\tr\,(g^{-1}
\hspace{-0.05cm}{_d\over^{dt}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g)^2
\nonumber} \newcommand{\no}{\noindent
\qqq
which stays constant during the evolution.
In the alternative way (\ref{alts}) to parametrize the solutions,
$\,h={1\over{4k}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\tr\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda^2\,$ and the symplectic
structure splits:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\Omega\ =\ \Omega_\ell\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}-\,\Omega_r\,,\quad\ \ {\rm where}
\quad\ \Omega_\ell\ =\ {_{i}\over^2}\,\tr\,
[\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}(g_\ell^{-1}dg_\ell)^2
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}-\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} d\lambda\, g_\ell^{-1}dg_\ell\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}]
\label{clsp}
\qqq
and $\Omega_r$ is given by the same formula with the subscript
$\ell$ replaced by $r$.
\vskip 0.3cm
There are two commuting actions of the group $G$
on ${\cal P}} \newcommand{\CQ}{{\cal Q}$: from the left $g(t)\mapsto g_0g(t)$ and from
the right $g(t)\mapsto g(t)g_0^{-1}$. Both preserve
the symplectic structure and the Hamiltonian $h$.
The vector fields corresponding to the left and right actions
of the infinitesimal generators $t^a\in{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$ are induced
by the functions
\begin{eqnarray}} \newcommand{\de}{\bar\partial
j^a&=&{_{ki}\over^2}\,\tr\,(t^a\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{_d\over^{dt}}g^{-1})\,=\,
{_1\over^2}\,\tr\,(t^a\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g_\ell\lambda\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g_\ell^{-1})\cr
\widetilde} \def\hat{\widehat j^a&=&{_{ki}\over^2}\,\tr\,(t^a\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
g^{-1}{_d\over^{dt}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g)\,=\,-\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{_1\over^2}\,\tr\,(t^a\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g_r
\lambda\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g_r^{-1}),
\nonumber} \newcommand{\no}{\noindent
\qqq
respectively. Note that, if we normalize $t^a$'s so that
$\tr\,(t^a t^b)={_1\over^2}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\delta^{ab}$, then
\begin{eqnarray}} \newcommand{\de}{\bar\partial
h\ =\ {_2\over^k}\, j^aj^a\ =\ {_2\over^k}\,\widetilde} \def\hat{\widehat j^a\widetilde} \def\hat{\widehat j^a
\nonumber} \newcommand{\no}{\noindent
\qqq
(summation convention!). The symplectic form $\Omega_\ell$
gives, for $\lambda$ fixed, the canonical symplectic form
on the (co)adjoint orbit $\,\{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g_\ell\lambda g_\ell^{-1}\,|
\,g_\ell\in G\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\}\,$
passing through $\lambda$. The left action of the group is
$\,g_\ell\mapsto g_0g_\ell\,$ so that it coincides with the
(co)adjoint action on the orbit. As is well known, upon geometric
quantization of the coadjoint orbits for appropriate $\lambda$,
this action gives rise to irreducible representations of $G$.
\vskip 0.4cm
\subsection{The quantization}
The geodesic motion on a group is easy to quantize.
As the Hilbert space ${\cal H}} \newcommand{\CI}{{\cal I}$ one takes the space
$L^2(G,dg)$ of functions on $G$ square integrable
with respect to the normalized Haar measure $dg$.
The two commuting actions of $G$ in ${\cal H}} \newcommand{\CI}{{\cal I}$:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
f\,\mapsto\,{}^{^{h}}\hspace{-0.14cm}f\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}=\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} f(h^{-1}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\cdot\,)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},
\qquad f\,\mapsto\,f^{^{h}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}=\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} f(\,\cdot\,h)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},
\nonumber} \newcommand{\no}{\noindent
\qqq
give rise to the actions
\begin{eqnarray}} \newcommand{\de}{\bar\partial
J^af\ =\ {_1\over^i}\,{_d\over^{d\epsilon}}\bigg\vert_{_{\epsilon=0}}
{}^{^{\ee^{i\epsilon t^a}}}\hspace{-0.14cm}f\,,\qquad
\widetilde} \def\hat{\widehat J^af\ =\ {_1\over^i}\,{_d\over^{d\epsilon}}
\bigg\vert_{_{\epsilon=0}}f^{^{\hspace{0.07cm}\ee^{i\epsilon t^a}}}
\nonumber} \newcommand{\no}{\noindent
\qqq
of the infinitesimal generators $t^a$ of ${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$.
The commutation relations
\begin{eqnarray}} \newcommand{\de}{\bar\partial
[J^a,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} J^b]\ =\ i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} f^{abc}\, J^c\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\qquad
[\widetilde} \def\hat{\widehat J^a,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\widetilde} \def\hat{\widehat J^b]\ =\ i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} f^{abc}\,\widetilde} \def\hat{\widehat J^c\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},
\nonumber} \newcommand{\no}{\noindent
\qqq
reflect the relation $[t^a,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} t^b]=i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} f^{abc}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} t^c$
in the Lie algebra ${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$. The quantum Hamiltonian
\begin{eqnarray}} \newcommand{\de}{\bar\partial
H\ =\ {_2\over^k}\, J^aJ^a\ =\ {_2\over^k}\,\widetilde} \def\hat{\widehat J^a\widetilde} \def\hat{\widehat J^a
\nonumber} \newcommand{\no}{\noindent
\qqq
coincides with $-\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{2\over k}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ times the Laplace-Beltrami
operator on $G$ and is a positive self-adjoint operator.
\vskip 0.3cm
The irreducible representations $R$ of the compact Lie group $G$
are finite dimensional and are necessarily unitarizable
so that we may assume that they act in finite-dimensional
vector spaces $V_{_R}$ preserving their scalar product.
We shall denote by $g_{_R}$ and $X_{_R}$ the endomorphisms
of $V_{_R}$ representing $g\in G$ and $X\in{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$.
Up to isomorphism, the irreducible representation
of $G$ may be characterized by their {\bf highest weights}.
Let us recall what this means.
The complexified Lie algebra may be decomposed into
the eigenspaces of the adjoint action of its Cartan subalgebra
$\Ntt$ as
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}^\NC\ =\ \Ntt^\NC\oplus(\mathop{\oplus}\limits_{\alpha}\NC\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
e_{\alpha})
\qqq
where \,$[X,e_\alpha]\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}=\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\tr\,(\alpha\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} X)\, e_\alpha\,$
for all $X\in\Ntt$. The set of the roots $\alpha\in\Ntt$
may be divided into the positive roots and their negatives.
We shall normalize the invariant form $\tr$ on ${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$
so that the long roots have the length squared 2 (this
agrees with the normalization of the matrix trace for
${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}=su(N)$). The ``step generators'' $e_{\pm\alpha}$ may
be chosen so that $[e_{\alpha},e_{-\alpha}]$ is equal to the
coroot \,$\alpha^\vee\equiv{2\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\alpha\over\hbox{tr}\,\alpha^2}\,$
corresponding to $\alpha$. The elements $\lambda\in\Ntt$ such
that $\tr\,(\alpha^\vee\lambda)$ is integer for all roots
are called weights. A non-zero vector $v\in V_{_R}$
(unique up to normalization) is called a highest weight (HW)
vector if it is an eigen-vector of the action of the Cartan
algebra: $\,X_{_R}v=\tr\,(\lambda_{_R}X)\, v$ and if
$(e_\alpha)_{_R}v=0$ for all positive roots $\alpha$.
The element $\lambda_{_R}$ of the Cartan algebra, a weight,
is called the highest weight (HW) of the representation $R$
and it determines completely $R$. All weights $\lambda\in\Ntt$
such that $\,\tr\,(\alpha^\vee\lambda)$ is a non-negative integer
for each positive $\alpha$ appear as HW's of irreducible
representations\footnote{such $\lambda$ are usually called
dominant weights} of $G$. The representations R may be obtained
by the geometric quantization of the (co)adjoint orbit
passing through $\lambda_{_R}$. \,For the $su(2)$ Lie algebra
spanned by the Pauli matrices
$\sigma_i$, one usually takes $\sigma_3$ as the positive root
and the matrices $\sigma_\pm={_1\over^2}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}(\sigma_1\pm i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\sigma_2)$
as the corresponding step generators.
The HW's are of the form $j\sigma_3$ with $j=0,{_1\over^2},1,\dots$
called the {\bf spin} of the representation.
\vskip 0.3cm
With respect to the left-right action of $G\times G$,
the Hilbert space $L^2(G,dg)$ decomposes as
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{\cal H}} \newcommand{\CI}{{\cal I}\ \cong\ \mathop{\oplus}\limits_{R}\,V_{_R}\otimes V_{_{\overline{R}}}\,,
\label{decomp}
\qqq
where the (infinite) sum is over the (equivalence classes of)
irreducible representations of $G$ and $\overline{R}$
denotes the representation complex-conjugate to $R$,
i.e. $V_{_{\overline{R}}}=\overline{V}_{_R}$ and
$g_{_{\overline{R}}}=\overline{g}_{_R}$.
\,Recall that the complex conjugate vector space $\overline{V}$
is composed of the vectors $v\in V$, denoted for distinction
by $\overline{v}$, with the multiplication by scalars defined by
$\mu\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\overline{v}=\overline{\bar{\mu} v}$. A linear transformation
$A$ of $V$, when viewed as a transformation $\overline{A}$ of $\overline{V}$,
is still linear. \,The above factorization of the
Hilbert space reflects the classical splitting (\ref{clsp}).
Let $(g_{_R}^{ij})$ be the (unitary)
matrix of the endomorphism $g_{_R}$ with respect
to a fixed orthonormal bases $(e_{_R}^i)$ in $V_{_R}$.
The decomposition (\ref{decomp}) is given by the assignment
\begin{eqnarray}} \newcommand{\de}{\bar\partial
V_{_R}\otimes V_{_{\overline{R}}}\,\ni\,
e_{_R}^i\otimes\overline{e_{_R}^j}\ \ \ \mapsto\ \ \ d_{_R}^{\,{_1\over^2}}\,
\,\overline{g_{_R}^{ij}}\,\in\,L^2(G,dg)\,.
\label{hwq}
\qqq
The Schur orthogonality
relations
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\int_{_G}\overline{g_{_{R'}}^{ij}}\,\,g_{_R}^{rs}
\,\,dg\ =\ {_1\over^{d_{_R}}}\,\delta_{_{R'R}}\,\,\delta^{ir}\,
\delta^{js}
\nonumber} \newcommand{\no}{\noindent
\qqq
assure that this assignment preserves the scalar product.
The matrix elements $g^{ij}_{_R}$ span a dense subspace
in $L^2(G,dg)$.
\vskip 0.3cm
The function on $G$ invariant under the adjoint
action $g\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\mapsto\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} Ad_{g_0}(g)\equiv g_0\, g\, g_0^{-1}$
are called class functions. They are constant on the conjugacy
classes
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\CC_{\lambda}=\{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g_0\,\ee^{\m2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} \lambda/k}
g_0^{-1}\,\vert\,g_0\in G\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\}
\qqq
with $\lambda$ in the Cartan algebra $\Ntt$.
The characters $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\chi_{_R}(g) =\tr_{_{V_{_{_R}}}}g_{_R}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ of the
irreducible representations $R$ are class functions. The Schur
relations imply that
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\int_{_G}\overline{\chi_{_{R'}}(g)}\,\,\chi_{_R}(g)\,\, dg\
=\ \delta_{_{R' R}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}.
\nonumber} \newcommand{\no}{\noindent
\qqq
The class functions in $L^2(G,dg)$ form a closed
subspace and the characters $\chi_{_R}$ form
an orthonormal bases of it.
Note that under the isomorphism (\ref{decomp}),
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\overline{\chi_{_R}}\ \ \cong\ \ d_{_R}^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}-{_1\over^2}}\,
e_{_R}^i\otimes\overline{e_{_R}^i}
\label{chdec}
\qqq
(sum over $i$\,!).
\vskip 0.3cm
The Hamiltonian $H$ becomes diagonal in the decomposition
(\ref{decomp}) of the Hilbert space. It acts on
$V_{_R}\otimes V_{_{\overline{R}}}$
as the multiplication by ${2\over k}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} c_{_R}$ where $c_{_R}$
is the value of the quadratic Casimir $c=t^at^a$ in the
representation $R$. In terms of the HW's, $c_{_R}={_1\over^2}\,\tr\,
(\lambda_{_R}(\lambda_{_R}+2\rho))\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$, \,where $\rho$,
the {\bf Weyl vector},
is equal to half the sum of the positive roots.
The Hamiltonian generates a 1-parameter family of
unitary transformation $\ee^{it\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} H}$ describing the time
evolution of the quantum system. In the Euclidean spirit,
we shall be more interested, however, in the semigroup
of the thermal density matrices $\ee^{-\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\beta\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} H}$
obtained by the Wick rotation of time $\beta=it\geq 0$.
Their (heat) kernels are given by:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\ee^{-\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} \beta\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} H}(g_0,g_1)\ =\ \sum\limits_R\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} d_{_R}\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
\ee^{-{_2\over^k}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} \beta\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} c_{_R}}\,\,\chi_{_R}(g_0g_1^{-1})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}.
\nonumber} \newcommand{\no}{\noindent
\qqq
In particular, at $\beta=0$, we obtain a representation for
the delta-function concentrated at an alement $g_0\in G$:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\delta_{g_0}(g_1)\ =\ \sum\limits_R\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} d_{_R}\,
\chi_{_R}(g_0g_1^{-1})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}.
\nonumber} \newcommand{\no}{\noindent
\qqq
We shall also need below the delta-functions concentrated on
the conjugacy classes $\CC_{\lambda}$. They may be obtained
by smearing the delta-function $\delta_{g}$ over $\CC_\lambda\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
&&\delta_{_{\CC_{\lambda}}}\hspace{-0.07cm}(g_1)
\ =\ \int_{_G}\hspace{-0.05cm}
\delta_{g_0\,\ee^{\m2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda/k}g_0^{-1}}(g_1)\,\, dg_0
\ =\ \sum\limits_Rd_{_R}\int_{_G}\hspace{-0.05cm}
\chi_{_R}(g_0\,\ee^{\m2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda/k}g_0^{-1}g_1^{-1})\,\,
dg_0\cr
&&=\ \sum\limits_Rd_{_R}\int_{_G}\hspace{-0.05cm}
(g_0)^{^{ij}}_{_R}\,(\ee^{\m2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda/k})^{^{j\ell}}_{_R}
\,\overline{(g_0)^{^{n\ell}}_{_R}}\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\overline{(g_1)^{^{in}}_{_R}}\,\,dg
\ =\ \sum\limits_R\chi_{_R}(\ee^{\m2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda/k})\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
\overline{\chi_{_R}(g_1)}\,,
\nonumber} \newcommand{\no}{\noindent
\qqq
where we have used the Schur relations.
It follows from the correspondence (\ref{chdec})
that in the language of the isomorphism (\ref{decomp}),
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\delta_{_{\CC_{\lambda}}}\ \cong\ \sum\limits_R
{\chi_{_R}(1)^{-{_1\over^2}}}\,\,{\chi_{_R}(\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} 2\pi
i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda/k})}\,\,e^{^i}_{_R}\otimes\overline{e^{^i}_{_R}}\,.
\label{Ishcl}
\qqq
More exactly, $\delta_{_{\CC_{\lambda}}}$ is not a normalizable
state in ${\cal H}} \newcommand{\CI}{{\cal I}$ but it defines an antilinear functional
on a dense subspace in ${\cal H}} \newcommand{\CI}{{\cal I}$, e.g.\,\,the one of vectors
with a finite number of components in the decomposition
(\ref{decomp}).
\vskip 0.3cm
The delta-functions $\delta_{_{\CC_{\lambda}}}$ may be
used to disintegrate the Haar measure $dg$ into the
measures along the conjugacy classes and over the
set of different conjugacy classes:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
dg\,=\,{_1\over^{\vert T\vert}}\,
\vert\Pi(\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} 2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda/k})\vert^2\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
\delta_{_{\CC_\lambda}}(g)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\,d\lambda\, dg\,.
\qqq
We shall choose the measure $d\lambda$ such
that it corresponds to the normalized Haar measure on
the Cartan group $T\subset G$ under the exponential map
$\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda\mapsto\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} 2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda/k}$. Then
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\Pi(\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} 2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda/k})=\prod\limits_{\alpha>0}(\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\pi
i\,\tr\,(\alpha\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda)/k}-\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\ee^{-\pi i\,\tr\,(\alpha\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda)
/k})
\label{WD}
\qqq
is the so called {\bf Weyl denominator}. In particular, for class
functions constant on the conjugacy classes one obtains
the Weyl formula:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\int_{_G}\hspace{-0.07cm}f\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} dg\ =\
\int f(\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} 2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda/k})\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\vert\Pi(\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} 2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
\lambda/k})\vert^2\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} d\lambda\,,
\label{WDcl}
\qqq
where each conjugacy class should be represented
ones. We shall employ this representation of the
integral of class functions later.
\vskip 0.3cm
The Feynman-Kac formula allows to express the heat kernel
on the group as a path integral:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\ee^{-\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} \beta\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} H}(g_0,g_1)\ \ =\int\limits_{g:\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}[0,\beta]\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
\rightarrow\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} G\atop g(0)=g_0,\ g(\beta)=g_1}\hspace{-0.2cm}
\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} -S(g)}\,\,Dg\,,
\nonumber} \newcommand{\no}{\noindent
\qqq
where $Dg$ stands for the product of the Haar measures
$dg(t)$. The integral on the right hand side may be
given a rigorous meaning as the one with respect to
the Brownian bridge measure $dW_{g_0,g_1}(g)$ supported
by continuous paths in $G$.
The path integral may be also used to define the
thermal {\bf correlation function}
\begin{eqnarray}} \newcommand{\de}{\bar\partial
<\prod\limits_ng^{^{i_nj_n}}_{_{R_n}}
(t_n)>_{_{\hspace{-0.08cm}\beta}}\ \ \equiv\ \
{{\int\prod\limits_{n=1}^N g^{^{i_nj_n}}_{_{R_n}}(t_n)
\,\,\ee^{-S(g)}\,\, Dg}\over
{\int\ee^{-S(g)}\,\, Dg}}
\,,
\label{fkp}
\qqq
where on the right hand side one integrates over the
periodic paths $g:[0,\beta]\rightarrow G$. Upon ordering
the (Euclidean) times $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} t_1\leq\dots\leq t_N\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$,
the above path integral may be expressed in the operator
language:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
&&<\prod\limits_ng^{^{i_nj_n}}_{_{R_n}}
(t_n)>_{_{\hspace{-0.08cm}\beta}}\cr
&&\hspace{0.7cm}=\ {{\tr_{_{{\cal H}} \newcommand{\CI}{{\cal I}}}\,\left(
\ee^{-t_1{\cal H}} \newcommand{\CI}{{\cal I}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g^{^{i_1j_1}}_{_{R_1}}
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\ee^{-(t_2-t_1){\cal H}} \newcommand{\CI}{{\cal I}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g^{^{i_2j_2}}_{_{R_2}}
\,\cdots\,\ee^{-(t_{_N}-t_{_{N-1}}){\cal H}} \newcommand{\CI}{{\cal I}}
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g^{^{i_{_N}j_{_N}}}_{_{R_N}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\ee^{-(\beta-t_{_N}){\cal H}} \newcommand{\CI}{{\cal I}}
\right)}\over{\tr_{_{\cal H}} \newcommand{\CI}{{\cal I}}\ \ee^{-\beta\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} H}}}\,,
\hspace{0.7cm}
\label{corf}
\qqq
where the functions $g^{ij}_{_R}$ on $G$ are viewed
as the multiplication operators in $L^2(G,dg)$.
\vskip 0.3cm
The right hand side of Eq.\,\,(\ref{corf}) may be calculated
using harmonic analysis on $G$. Indeed, what is really needed
for such a computation are the matrix elements
\begin{eqnarray}} \newcommand{\de}{\bar\partial
(\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} e_{_{R_1}}^{^{i_1}}\otimes
\overline{e_{_{R_1}}^{^{j_1}}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},
\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g_{_R}^{^{ij}}\,\, e_{_{R_2}}^{^{i_2}}
\otimes\overline{e_{_{R_2}}^{^{j_2}}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch})
\ =\ d_{_{R_1}}^{_1\over^2} d_{_{R_2}}^{{_1\over^2}}
\int_{_G}\hspace{-0.07cm}
\overline{g_{_{R_2}}^{^{i_2j_2}}}\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g_{_{R_1}}^{^{i_1j_1}}
\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g_{_R}^{^{ij}}\,\, dg
\qqq
encoding the decomposition of the tensor product of
the irreducible representations
\begin{eqnarray}} \newcommand{\de}{\bar\partial
V_{_R}\otimes V_{_{R_1}}\ \cong\ \mathop{\oplus}\limits_{R_2}
M_{_{R_1R}}^{^{R_2}}\otimes V_{_{R_2}}\,.
\label{tpd}
\qqq
In particular, the dimensions $N_{_{R_1R}}^{^{R_2}}$
of the multiplicity spaces $M_{_{R_1R}}^{^{R_2}}$
may be obtained from the traces of the above matrix elements:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
N_{_{R_1R}}^{^{R_2}}&=&\int_{_G}\hspace{-0.07cm}
\overline{\chi_{_{R_2}}(g)}\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\chi_{_{R_1}}(g)\,
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\chi_{_{R}}(g)\,\,dg\,.
\nonumber} \newcommand{\no}{\noindent
\qqq
The finite combinations with integer coefficients
$\sum n_i\chi_{_{R_i}}$ of characters of irreducible
representations form a subring ${\cal R}} \newcommand{\CS}{{\cal S}_{_G}$ in the commutative
ring of class functions. The identity
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\chi_{_R}\,\chi_{_{R_1}}\ =\ \sum\limits_{R_2}
N_{_{R\, R_1}}^{^{R_2}}\,\chi_{_{R_2}}
\label{strf}
\qqq
shows that the integers $N_{_{R_1R}}^{^{R_2}}
=N_{_{R\, R_1}}^{^{R_2}}$
play the role of structure constants of this ring.
\vskip 0.3cm
One can define a version of the correlation functions
by replacing the integral
over the periodic paths $g:[0,\beta]\rightarrow G$
in Eq.\,\,(\ref{fkp})
by the one over the paths constraint
to fixed conjugacy classes on the boundary
of the interval:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{{\int\limits_{g:[0,\beta]\rightarrow G}
\prod\limits_{n=1}^N g^{^{i_nj_n}}_{_{R_n}}(t_n)
\,\,\delta_{_{\CC_{\lambda_1}}}
\hspace{-0.07cm}(g(0))\,\,\delta_{_{\CC_{\lambda_2}}}
\hspace{-0.07cm}(g(\beta))\,\,\ee^{-S(g)}\,\,Dg}
\over
{\int\limits_{g:[0,\beta]\rightarrow G}
\delta_{_{\CC_{\lambda_1}}}
\hspace{-0.07cm}(g(0))\,\,\delta_{_{\CC_{\lambda_2}}}
\hspace{-0.07cm}(g(\beta))\,\,\ee^{-S(g)}\,\,Dg}}
\ \equiv\ <\prod\limits_ng^{^{i_nj_n}}(t_n)
>_{_{\hspace{-0.08cm}\beta,
\lambda_1\lambda_2}}\,.\hspace{0.7cm}
\qqq
For $G=SU(2)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}=\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} x_0+i x_i\sigma_i\,\vert\,
x_0^2+x_i^2=1\}\cong S^3$, the conjugacy classes
are the 2-spheres with fixed $x_0$ so that
one integrates over the paths as in Fig.\,\,1.
\leavevmode\epsffile[-110 -20 310 193]{fig1.eps}
The above functional integral may be rewritten
in the operator language as
\begin{eqnarray}} \newcommand{\de}{\bar\partial
<\prod\limits_ng^{^{i_nj_n}}(t_n)>_{_{\hspace{-0.08cm}\beta,
\lambda_1\lambda_2}}
\,=\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{{\Big(\delta_{_{\CC_{\lambda_1}}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
\ee^{-t_1{\cal H}} \newcommand{\CI}{{\cal I}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g^{^{i_1j_1}}_{_{R_1}}
\,\cdots\,\ee^{-(t_{_N}-t_{_{N-1}}){\cal H}} \newcommand{\CI}{{\cal I}}
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g^{^{i_{_N}j_{_N}}}_{_{R_N}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\ee^{-(\beta-t_{N}){\cal H}} \newcommand{\CI}{{\cal I}}\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
\delta_{_{\CC_{\lambda_2}}}\Big)}\over
{\Big(\delta_{_{\CC_{\lambda_1}}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
\ee^{-\beta\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\cal H}} \newcommand{\CI}{{\cal I}}\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\delta_{_{\CC_{\lambda_2}}}\Big)}}
\,.\hspace{0.8cm}
\label{bdcf}
\qqq
Although $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\delta_{_{\CC_{\lambda}}}$ are generalized
functions rather than normalizable states in $L^2(G,dg)$,
the matrix elements on the right hand side
are finite and may again be computed by harmonic analysis
on $G$.
\vskip 0.3cm
We shall encounter field-theoretical generalization
of the above quantum-mechanical constructions below.
\vskip 0.6cm
\nsection{The WZW action}
\subsection{Two-dimensional sigma models}
The two-dimensional sigma models describe field theories
with fields mapping a surface $\Sigma$
to a target manifold $M$, both equipped with metric
structures. Such field configurations represent
evolution of a string in the target $M$ with
$\Sigma$ being the string world sheet.
The (Euclidean) action functional of the field
configuration $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} X:\Sigma\rightarrow M\,$ is
\begin{eqnarray}} \newcommand{\de}{\bar\partial
S^\gamma(X)\ =\ {_1\over^{4\pi}}\int_{_\Sigma}\hspace{-0.1cm}
\gamma_{\mu\nu}(X)\,\partial} \newcommand{\ee}{{\rm e}_\alpha X^\mu\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\partial} \newcommand{\ee}{{\rm e}_\beta X^\nu
\eta^{\alpha\beta}\,\sqrt{\eta}\,,
\qqq
where $\gamma_{\mu\nu}$ is the Riemannian
metric on $M$, $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\eta_{\alpha\beta}$ the one
on $\Sigma$ and $\sqrt{\eta}\equiv\sqrt{\det{
\eta_{\alpha\beta}}}$ is
the Riemannian volume density on $\Sigma$.
In particular, if $M=\NR^n$ with the standard metric,
we obtain the quadratic action of the free field
on a two-dimensional surface leading to linear
classical equations. The general case, however, results
in a non-linear classical theory.
\vskip 0.3cm
The term $S^\gamma$ does not change under the local
rescalings $\eta_{\alpha\beta}\mapsto\ee^{2\sigma}
\eta_{\alpha\beta}$ of the metric on $\Sigma$
i.e. it possesses two-dimensional conformal invariance.
For oriented $\Sigma$, conformal classes of the
metric are in one to one correspondence with
complex structures on $\Sigma$ such that $\eta_{zz}
=\eta_{\bar z\bar z}=0$ in the holomorphic coordinates
and that the latter preserve the orientation.
The action $S^\gamma$ may be written using
explicitly only the complex structure of $\Sigma\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
S^\gamma(X)\ =\ {_i\over^{2\pi}}\int_{_\Sigma}\hspace{-0.1cm}
\gamma_{\mu\nu}(X)\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\partial} \newcommand{\ee}{{\rm e} X^\mu\,\de X^\nu\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},
\qqq
where $\partial} \newcommand{\ee}{{\rm e}=dz\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\partial} \newcommand{\ee}{{\rm e}_z$ and $\de=d\bar z\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\partial} \newcommand{\ee}{{\rm e}_{\bar z}$.
One-dimensional complex manifolds are called Riemann surfaces.
It follows that the action $S^\gamma(X)$ may be defined
on such surfaces.
\vskip 0.3cm
To the $S^\gamma$ term, one may add the expression
\begin{eqnarray}} \newcommand{\de}{\bar\partial
S^\beta(X)\ =\ {_1\over^{4\pi i}}
\int_{_\Sigma}\hspace{-0.1cm}\beta_{\mu\nu}(X)\,
\partial} \newcommand{\ee}{{\rm e}_\alpha X^\mu\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\partial} \newcommand{\ee}{{\rm e}_\beta X^\nu \epsilon^{\alpha\beta}
\qqq
where $\beta_{\mu\nu}=-\beta_{\nu\mu}$ are the coefficients
of a 2-form $\beta$ on $M$. Geometrically, $S^\beta$ is
proportional to the integral of the pullback of $\beta$
by $X$:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
S^\beta(X)\ =\ {_1\over^{4\pi i}}
\int_{_\Sigma}\hspace{-0.1cm}X^*\beta\,.
\qqq
The imaginary coefficient is required by the unitarity
of the theory after the Wick rotation to the Minkowski
signature. The term $S^\beta$ does not use the metric
on $\Sigma$ but only the orientation and is often called
a topological term. Hence the classical two-dimensional
conformal invariance of the model with the action
$S=S^\gamma+S^\beta$.
\vskip 0.3cm
On the quantum level, the sigma model requires a renormalization
which often imposes the addition to the action of further terms
\begin{eqnarray}} \newcommand{\de}{\bar\partial
S^{tach}(X)\ =\ {_1\over^{2\pi}}
\int_{_\Sigma}\hspace{-0.09cm}{\cal T}} \newcommand{\CU}{{\cal U}(X)\,\sqrt{\eta}\quad\ \
{\rm and}\ \ \quad
S^{dil}(X)\ =\ {_1\over^{2\pi}}\int_{_\Sigma}\hspace{-0.08cm}
{\cal D}} \newcommand{\CE}{{\cal E}(X)\,\,r\,\sqrt{\eta}\,,
\qqq
where
where ${\cal T}} \newcommand{\CU}{{\cal U}$ and ${\cal D}} \newcommand{\CE}{{\cal E}$ are functions on $M$ called tachyonic
and dilatonic potentials, respectively, and $r$ is the scalar
curvature of the metric $\eta_{\alpha\beta}$.
The renormalization breaks the conformal
invariance (note that $S^{tach}$ and $S^{dil}$ are
not conformal invariant). We shall be interested, however,
in the case of the WZW model \cite{4}, an example of a CFT,
where the classical conformal invariance is (almost)
not broken on the quantum level.
\vskip 0.3cm
The WZW model is the two-dimensional counterpart of the
particle on a group and may be thought of as describing
the movement of a string on a group manifold $M=G$ equipped
with the invariant metric $\gamma$ described before.
We then have for $g:\Sigma\rightarrow G$,
\begin{eqnarray}} \newcommand{\de}{\bar\partial
S^\gamma(g)\ =\ {_k\over^{4\pi i}}\int_{_\Sigma}\hspace{-0.09cm}
\tr\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}(g^{-1}\partial} \newcommand{\ee}{{\rm e} g)(g^{-1}\de g)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},
\label{sgam}
\qqq
where $k$ is a positive constant. The quantization
of a model with such an action leads, however, to
a theory without conformal invariance. To restore
the latter, one adds to the $S^\gamma$ term, following Witten
\cite{4}, a topological term, the so called
Wess-Zumino (WZ) term $S^{WZ}$. In the first approximation,
$S^{WZ}=k\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} S^\beta$ where $\beta$ is a 2-form on $G$ such that
$d\beta$ is equal to the canonical 3-form $\chi\equiv{1\over 3}
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\tr\,(g^{-1}dg)^3$ on $G$. If the group $G$ is abelian,
such a description is indeed possible and the overall
action is a simple version of the free field one.
In the non-abelian case, however, the difficulty comes from
the fact that the 3-form $\chi$ is closed but not globally
exact so that the forms $\beta$ exist only locally
and are defined only up to closed 2-forms. Hence
the definition of the WZ term of the action requires
a more refined discussion.
\vskip 0.4cm
\subsection{Particle in the field of a magnetic monopole}
It may be useful to recall a simpler situation where
one is confronted with a similar problem. Suppose that we
want to define the contribution $S^{Dir}(x)$ to the action
of a mechanical particle of the term
$$e\int_{_0}^{^T}\hspace{-0.07cm}A_\nu(x)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} {dx^\nu\over dt
\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} dt\ =\ e\int x^*A$$ describing
the coupling to the abelian gauge field $A=A_\nu dx^\nu$
with the field strength $F_{\nu\lambda}={_1\over^2}(\partial} \newcommand{\ee}{{\rm e}_\nu A_\lambda
-\partial} \newcommand{\ee}{{\rm e}_\lambda A_\nu)$, or in the language of differential forms,
with $F=dA\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$, \hspace{0.025cm}} \newcommand{\ch}{{\rm ch} where $F=F_{\nu\lambda}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} dx^\nu dx^\lambda$.
The constant $\,e\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ stands for the electric charge
of the particle.
For concreteness, suppose that $F_{\mu\nu}$ corresponds to
the magnetic field of a monopole of magnetic charge $\mu$
placed at the origin of $\NR^3$:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
F_{\nu\lambda}\,=\,{_1\over^2}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} \mu\,\epsilon_{\nu\lambda\kappa}\,
{{x^\kappa}\over{\vert x\vert^3}}\,.
\nonumber} \newcommand{\no}{\noindent
\qqq
There is no global 1-form $A$ on $\NR^3$ without the origin
such that $dA=F$. For a closed trajectory $t\mapsto x(t)$,
however, we may pose
\begin{eqnarray}} \newcommand{\de}{\bar\partial
S^{Dir}(x)\ =\ e\int_{_D}\hspace{-0.08cm}\widetilde} \def\hat{\widehat x^*F\,,
\qqq
where $\widetilde} \def\hat{\widehat x$ is a map of a disc $D$ into
$\NR^3\setminus\{0\}$ coinciding on the boundary of the disc
with $x$. For two different extensions $\widetilde} \def\hat{\widehat x$, however,
the above prescription may give different results.
Their difference may be written as the integral
\begin{eqnarray}} \newcommand{\de}{\bar\partial
e\int_{_{S^2}}\hspace{-0.08cm}\widetilde} \def\hat{\widehat x^*F
\label{inF}
\qqq
over the 2-sphere $S^2$ obtained by gluing the two disc $D$, one
with the inverted orientation, along the boundary and for the map
$\,\widetilde} \def\hat{\widehat x:S^2\rightarrow\NR^3\setminus\{0\}\,$ glued from
the two extensions of $x$ to the respective discs, see Fig.\,\,2.
\leavevmode\epsffile[-150 -15 270 156]{fig2.eps}
\vskip 0.1cm
The ambiguities (\ref{inF}) are the periods of the closed
form $F$ over the cycles of the 2$^{\rm nd}$ integer homology
$\,H_2(\NR^3\setminus\{0\})=\NZ$. They take discrete values which
are multiples of $4\pi\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} e\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} \mu$. The latter value
is obtained for the unit sphere in $\NR^3$,
a generator of $H_2(\NR^3\setminus\{0\})$.
The discrete ambiguities are acceptable in classical mechanics
where one studies the extrema of the action. In quantum mechanics,
however, we have to give sense to the Feynman amplitudes
$\,\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} S^{Dir}(x)}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$, hence only the ambiguities in the
action with values in $2\pi\NZ$ are admissible. Demanding that
the quantum-mechanical amplitudes be unambiguously defined
reproduces this way the Dirac quantization condition
$\,e\mu\in{_1\over^2}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\NZ$.
\vskip 0.3cm
For open trajectories $[0,T]\ni t\mapsto x(t)$,
the amplitudes $\,\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} e\int x^*A}$ may not, in general,
be unambiguously assigned numerical values. They may be only defined
as maps between the fibers ${\cal L}} \newcommand{\CM}{{\cal M}_{x(0)}$ and ${\cal L}} \newcommand{\CM}{{\cal M}_{x(T)}$
of a line bundle ${\cal L}} \newcommand{\CM}{{\cal M}$. Geometrically, they give the parallel
transport in the bundle corresponding to a $U(1)$-connection
with the curvature form $F$. We shall recover the analogous
situation below when discussing how to give meaning to
the WZ term in the action of the WZW model.
\vskip 0.4cm
\subsection{Wess-Zumino action on surfaces without boundary}
Let us first consider the case of compact Riemann
surfaces without boundary.
\leavevmode\epsffile[-100 5 320 160]{fig3.eps}
\noindent Topologically, such surfaces
are characterized by the genus $g_{_\Sigma}$ equal
to the number of handles of the surface. They may be viewed
as world sheets of a closed string created
from the vacuum, undergoing in the evolution $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g_{_\Sigma}$
splittings and recombinations and finally disappearing
into the vacuum, see Fig.\,\,3 where a surface of genus 2
was represented.
At genus zero, there is only one (up to diffeomorphisms)
Riemann
surface, the Riemann sphere $\NC P^1=\NC\cup\{\infty\}$,
see Fig.\,\,4(a).
\leavevmode\epsffile[-60 -20 347 175]{fig4.eps}
\noindent At genus one, there is a complex one-parameter family
of Riemann surfaces: the complex tori
$\,{T}_\tau=\NC/(\NZ+\tau\NZ)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ with $\tau$
in the upper half plane $H^+\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ of the complex numbers
such that $\,{\rm Im}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\tau>0$, see Fig.\,\,4(b).
The tori ${T}_\tau$ and ${T}_{\tau'}$, where $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\tau'
={a+b\tau\over c+d\tau}\,$ for $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}({_a\atop^c}
\,{_b\atop^d})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ in the modular group $SL(2,\NZ)$,
may be identified by the map $\,z\mapsto z'=(c\tau+d)^{-1}z\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$.
The space of the diffeomorphism classes (i.e.\,\,the {\bf moduli
space}) of genus one Riemann surfaces is equal to $H^+/SL(2,\NZ)$
and has complex dimension 1. For higher genera,
the moduli spaces of Riemann surfaces have complex
dimension $3(g_{_\Sigma}-1)$.
\vskip 0.3cm
Let us return to the discussion of the action of the WZW model.
Assume that $G$ is connected and simply connected
and that $\Sigma$ is a compact Riemann surface without boundary.
Following \cite{4} and mimicking the trick used for a particle
in a monopole field, one may extend the field $g:\Sigma
\rightarrow G$ to a map $\widetilde} \def\hat{\widehat g:B\rightarrow G$
of a 3-manifold $B$ such that $\partial} \newcommand{\ee}{{\rm e} B=\Sigma$ and set:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
S^{^{WZ}}(g)\ =\ {_k\over^{4\pi i}}\int_{_B}\hspace{-0.07cm}
\widetilde} \def\hat{\widehat g^*\chi\,.
\label{28}
\qqq
By the Stokes formula, this expression coincides with
$k\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} S^\beta(g)$ whenever the image of $\widetilde} \def\hat{\widehat g$ is contained
in the domain of definition of a 2-form $\beta$ such that
$d\beta=\chi$, but it makes sense in the general case.
The price is that the result depends on the extension
$\widetilde} \def\hat{\widehat g$ of the field $g$. The ambiguities have the form
of the integrals
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{_k\over^{4\pi i}}\int_{_{\widetilde} \def\hat{\widehat B}}\hspace{-0.07cm}
\widetilde} \def\hat{\widehat g^*\chi
\label{per}
\qqq
over 3-manifolds $\widetilde} \def\hat{\widehat B$ without boundary with
$\widetilde} \def\hat{\widehat g:\widetilde} \def\hat{\widehat B\rightarrow G$, see Fig.\,\,5.
\leavevmode\epsffile[-35 -20 310 150]{fig5.eps}
\noindent They are proportional
to the periods of the 3-form $\chi$ over the integer
homology $H_3(G)$. Such discrete contributions do not
effect the classical equations of motion $\delta S=0$.
In quantum mechanics, however, where we deal
with the Feynman amplitudes\footnote{there is no $i$
in front of $S$ since we work with the Euclidean
action} $\,\ee^{-S(g)}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$, \hspace{0.025cm}} \newcommand{\ch}{{\rm ch} only ambiguities in
$2\pi i\NZ$ are allowed. Hence we have to find conditions
under which the periods (\ref{per}) lie in $2\pi i\NZ$.
\vskip 0.3cm
Recall that we normalized the invariant form $\tr$ on the Lie
algebra ${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$ of $G$ so that the long roots in $\Ntt$ have
length squared 2. When $\,G=SU(2)\cong\{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} x\in\NR^4\,\vert\,
\vert x\vert^2=1\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$, \,the 3-form $\chi$ equals
then to $4$ times the volume
form of the unit 3-sphere. Since the volume of the latter
is equal to $2\pi^2$, we infer that
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{_1\over^{4\pi i}}
\int_{_{SU(2)}}\hspace{-0.06cm}\chi\ =\ -2\pi i\,.
\qqq
For the other simple, simply connected groups, the roots
$\alpha$ determine the sub-algebras $su(2)_{_\alpha}\subset{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$,
obtained by identifying the corresponding coroot $\alpha^\vee=
{2\,\alpha\over{\hbox{tr}\,\alpha^2}}$ and the step generators
$e_{\pm\alpha}$ with the Pauli matrices $\sigma_3$ and
$\sigma_{\pm}$, respectively. By exponentiation, we obtain
the $SU(2)_{_\alpha}$ subgroups of $G$. Clearly,
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{_1\over^{4\pi i}}
\int_{_{SU(2)_{_\alpha}}}\hspace{-0.06cm}\chi\
=\ -{_{4\pi i}\over^{\hbox{tr}\,\alpha^2}}\,.
\qqq
The ratio $2\over\tr\,\alpha^2$ is equal to 1 for long roots
and is a positive integer for the others. It appears
that any of the subgroups $SU(2)_{_\alpha}\cong S^3\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$
for $\alpha$ a long root generates $H_3(G)=\NZ$.
Thus the unambiguous definition of the amplitudes
$\ee^{-S^{WZ}(g)}$ requires that the coupling
constant $k$, called the {\bf level} of the model, be a
(positive) integer, in the analogy to the Dirac quantization
of the magnetic charge.
\vskip 0.3cm
It is easy to see that, although the action $S^{WZ}(g)$
cannot be expressed, in general, as a local integral
over $\Sigma$, the variation of $S^{^{WZ}}$ has such a form:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\delta S^{^{WZ}}(g)\ =\ {_k\over^{4\pi i}}\int_{_\Sigma}
\hspace{-0.09cm}\tr\,(g^{-1}\delta g)(g^{-1}dg)^2\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}.
\label{vWZ}
\qqq
The above formula is a special case of the general, very useful,
geometric identity: $\,\delta\hspace{-0.05cm}\int
\hspace{-0.05cm}f^*\alpha=\int\hspace{-0.05cm}{\cal L}} \newcommand{\CM}{{\cal M}_{\delta f}
\alpha\,$, \,where ${\cal L}} \newcommand{\CM}{{\cal M}_{\cal X}} \newcommand{\CY}{{\cal Y}$ is the Lie derivative.
Applied to $f=g$ and $\alpha=\chi$ it gives,
in conjunction with the Stokes formula,
the above relation. \,It is also important to note the behavior
of $S^{^{WZ}}$ under the point-wise multiplication of fields,
a basic property of the WZ term:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
S^{^{WZ}}(g_1g_2)\ =\ S^{^{WZ}}(g_1)\,+\, S^{^{WZ}}(g_2)\,+\,
W(g_1,g_2)\,,
\label{pwWZ}
\qqq
where
\begin{eqnarray}} \newcommand{\de}{\bar\partial
W(g_1,g_2)\,=\,
{_k\over^{4\pi i}}\int_{_{\Sigma}}\hspace{-0.1cm}\tr
\,(g_1^{-1}dg_1)(g_2dg_2^{-1})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}.
\label{WWZ}
\qqq
The relation follows easily from the definition (\ref{28}),
again by applying the Stokes formula.
\vskip 0.3cm
The complete action of the WZW model on a closed Riemann surface
$\Sigma$ is the sum of the $\gamma$- and the WZ-terms with
the same coupling constant $k$: $\,S(g)=S^\gamma(g)+S^{^{WZ}}(g)$.
Since $S^\gamma$ is unambiguous, it has the same ambiguities
as $S^{WZ}$, requiring that $k$ be
a (positive) integer. The relations (\ref{vWZ}) and (\ref{pwWZ})
get also contributions from $S^\gamma$ and become:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
&&\delta S(g)\ =\ -\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{_k\over^{2\pi i}}\int_{_\Sigma}
\hspace{-0.09cm}\tr\,(g^{-1}\delta g)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\partial} \newcommand{\ee}{{\rm e}(g^{-1}\de g)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},
\label{vWZW}\\\cr
&&S(g_1g_2)\ =\ S(g_1)\,+\, S(g_2)\,+\,
{_k\over^{2\pi i}}\int_{_{\Sigma}}\hspace{-0.1cm}\tr
\,(g_1^{-1}\de g_1)(g_2\partial} \newcommand{\ee}{{\rm e} g_2^{-1})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}.
\label{pwWZW}
\qqq
The last relation is often called the Polyakov-Wiegmann formula.
From Eq.\,\,(\ref{vWZW}), we obtain the classical equations
of motion
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\partial} \newcommand{\ee}{{\rm e}(g^{-1}\de g)\,=\,0\,\qquad{\rm or, \,\,equivalently,}
\qquad\de(g\partial} \newcommand{\ee}{{\rm e} g^{-1})\,=\,0\,.
\label{cleq}
\qqq
They have few solutions with values in $G$ (this would not be
the case if we considered $\Sigma$ with a Minkowski metric).
In all the above formulae, however, we could have taken fields
$g$ with values in the complexified group $G^\NC$. For such
fields, the general local solutions of Eqs.\,\,(\ref{cleq})
have the form
\begin{eqnarray}} \newcommand{\de}{\bar\partial
g(z,\bar z)\ =\ g_\ell(z)\,g_r(\bar z)^{-1}
\label{lcls}
\qqq
where $g_\ell$ ($g_r$) are local holomorphic (anti-holomorphic)
maps with values in $G^\NC$. Thus Eqs.\,\,(\ref{cleq})
constitute a non-linear generalization of the Laplace equation
in two dimensions whose solutions are harmonic functions
which are, locally, sums of holomorphic and anti-holomorphic
ones. In particular, multiplying a solution (\ref{lcls}) by
a holomorphic map into $G^\NC$ on the left and by an anti-holomorphic
one on the right we obtain another solution. Similarly, composing
a solution with a local holomorphic map or inverting it in $G^\NC$
after composition with a local anti-holomorphic map of $\Sigma$
one produces new solutions. Hence a rich symmetry structure
of the classical theory. This structure will be preserved
by the quantization leading to the current and Virasoro algebra
symmetries of the quantum WZW model.
\vskip 0.4cm
\subsection{Wess-Zumino action on surfaces with boundary}
What if the surface $\Sigma$ has a boundary? Of course
only the WZ term in the action causes problems
due to its non-local character. The term $S^\gamma$
is defined unambiguously for any compact surface. It will
be convenient
to represent $\Sigma$ as $\Sigma'\setminus(\mathop{\cup}
\limits_nD_{n}^{^{^{{\hspace{-0.21cm}}o}}})$
where $D_{n}$ are disjoint unit discs $\{\,z\,\vert\,\vert z\vert\leq 1\}$
embedded in a closed surface $\Sigma'$ without boundary,
see Fig.6.
\leavevmode\epsffile[-80 -20 337 170]{fig6.eps}
\noindent Note that the boundaries of $\Sigma$ are then
naturally parametrized by the unit circles. One way to proceed
in the presence of boundaries is to extend the field
$g:\Sigma\rightarrow G$ to a map $g':\Sigma'\rightarrow G$
and to consider the action $\,S^{{WZ}}_{_{\Sigma'}}(g')\,$
pertaining to the surface $\Sigma'$, as stressed
by a subscript. We are again confronted with the question
as to how the action on $\Sigma'$ depends on the extension
of the field. The answer is easy to work out.
If $g''=g'h$ is another extension of $g$
then, by Eq.\,\,(\ref{pwWZ}),
\begin{eqnarray}} \newcommand{\de}{\bar\partial
S^{^{WZ}}_{{\Sigma'}}(g'')\ =\ S^{^{WZ}}_{{\Sigma'}}(g')\,
+\,S^{^{WZ}}_{{\Sigma'}}(h)\,+\,W_{{\Sigma'}}(g',h)\,.
\qqq
It will be convenient to localize
the changes in the discs $D_{n}$
by rewriting the last formula as
\begin{eqnarray}} \newcommand{\de}{\bar\partial
S^{^{WZ}}_{{\Sigma'}}(g'')\ =\ S^{^{WZ}}_{{\Sigma'}}(g')\,+
\sum\limits_n\left(S^{^{WZ}}_{{S^2}}
(h_n)\,+\,W_{_{D_{_n}}}(g',h)\right),
\label{trpro}
\qqq
where $h_n$, mapping spheres (compactified planes) $S^2_n$
to $G$, extend the maps $h\vert_{_{D_{_n}}}$ by unity and
$W_{_{D_{_n}}}$ are as in Eq.\,\,(\ref{WWZ}) but with
the integration restricted to $D_{n}$. To account for
the change (\ref{trpro}), we shall define
the following equivalence relation between the pairs $(g',z)$
where $g':D\rightarrow G$ and $z\in\NC$\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} :
\begin{eqnarray}} \newcommand{\de}{\bar\partial
(g',\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} z)\ \sim\ (g'h,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} z\,\ee^{-S^{{WZ}}_{S^2}(h)-
W_{_{D}}(g',h)})
\nonumber} \newcommand{\no}{\noindent
\qqq
for $h:S^2\rightarrow G$ equal to unity outside the
unit disc $D\subset S^2$. The set of equivalence classes
forms a complex line bundle ${\cal L}} \newcommand{\CM}{{\cal M}$ over the loop group $LG$
of the boundary values of the maps $g'$. Comparing
to Eq.\,\,(\ref{trpro}), we infer that for
$g:\Sigma\rightarrow G$ with $\Sigma$ as above,
the amplitude $\ee^{-S^{{WZ}}_{{\Sigma}}(g)}$ makes sense
as the element of a tensor product of the line bundles ${\cal L}} \newcommand{\CM}{{\cal M}$,
one for each boundary component of $\Sigma\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$,
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\ee^{-S^{WZ}_{{\Sigma}}(g)}\ \in\ \mathop{\otimes}\limits_n
{\cal L}} \newcommand{\CM}{{\cal M}_{_{g|_{_{\partial} \newcommand{\ee}{{\rm e} D_{_n}}}}}\,,
\nonumber} \newcommand{\no}{\noindent
\qqq
where ${\cal L}} \newcommand{\CM}{{\cal M}_{_h}$ denotes the fiber of ${\cal L}} \newcommand{\CM}{{\cal M}$ over the loop
$h\in LG$. Hence the WZ amplitudes
$\,\ee^{-S^{WZ}_{{\Sigma}}(g)}\,$
take values in line bundles instead of having numerical values,
exactly as for the amplitudes giving the parallel transport
in a $U(1)$-gauge field mentioned before.
\vskip 0.3cm
The line bundle ${\cal L}} \newcommand{\CM}{{\cal M}$ is an interesting object.
It carries a hermitian structure given by the absolute value
of $z$. The fibers of ${\cal L}} \newcommand{\CM}{{\cal M}$ over $g$ and $\check g$
where $\check g$ ia a reversed loop, $\check g(\ee^{i\varphi})
=g(\ee^{-i\varphi})$, may be naturally
paired so that, for $g:\widetilde} \def\hat{\widehat\Sigma
\rightarrow G$, where $\widetilde} \def\hat{\widehat\Sigma$ is obtained from two
surfaces $\Sigma$ and $\Sigma'$ by gluing them along some
boundary components, see Fig.\,\,7,
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\langle\,\ee^{-S^{WZ}_{_\Sigma}(g|_{_\Sigma})},\,
\ee^{-S^{WZ}_{_{\Sigma'}}
(g|_{_{\Sigma'}})}\,\rangle\ =\ \ee^{-S_{_{\widetilde} \def\hat{\widehat\Sigma}}(g)}\,.
\label{glue}
\qqq
\leavevmode\epsffile[-100 -20 317 180]{fig7.eps}
\noindent${\cal L}} \newcommand{\CM}{{\cal M}\,$ may be also equipped with a product structure
such that
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\ee^{-S_{_\Sigma}(g_1)}\cdot\ee^{-S_{_\Sigma}(g_2)}\ =\
\ee^{-S_{_\Sigma}(g_1g_2)}\,\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} W_{_\Sigma}(g_1,g_2)}\,,
\label{prKM}
\qqq
compare to Eq.\,\,(\ref{pwWZ}). Under the product,
the elements of unit length in ${\cal L}} \newcommand{\CM}{{\cal M}$
form a group $\hat{G}$ which is a {\bf central extension} of the
loop group $LG$ by the circle group $U(1)$:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
1\rightarrow U(1)\rightarrow\hat{G}\rightarrow LG\rightarrow 1\,.
\qqq
The second arrow sends $\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} i\varphi}$ to the equivalence
class of $(1,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} i\varphi})$. The extensions for $k>1$
are powers of the universal one corresponding to $k=1$. On the
infinitesimal level, one obtains the central extensions
of the loop algebra $L{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$ of the maps of the circle
$S^1$ to the Lie algebra ${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$ by the real line:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
0\rightarrow\NR\rightarrow\hat{{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}}\rightarrow L{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}\rightarrow 0\,.
\label{exs}
\qqq
The Lie algebra $\hat{{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}}$, called the {\bf current} or
the affine Kac-Moody {\bf algebra}, may be described explicitly
in terms of the complexified generators $t^a_n$ corresponding
to the loops $t^a\ee^{in\varphi}$ in $L{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}^\NC$ and the central
element $K$ satisfying the commutation relations:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
[t^a_n,t^b_m]\,=\, i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} f^{abc}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} t^c_{n+m}\,+\,{_1\over^2}
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} K\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} n\,\delta^{ab}\delta_{n+m,0}\,.
\label{KM}
\qqq
The algebra $\hat{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$ is the same
for all levels $k$ but the central element
$K\in\hat{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$ is the image of $k\in\NR$ under the second arrow
in the exact sequences (\ref{exs}). Note that the generators $t^a_0$
span a subalgebra ${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}\subset\hat{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$. As we shall see, the group
$\hat G$ and the algebra $\hat{{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}}$ play in the WZW theory a similar
role to that of $G$ and ${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$ for the particle on the group.
\vskip 0.4cm
\subsection{Coupling to gauge field}
We may couple the WZW model to the gauge field
$i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A\equiv i(A^{10}+A^{01})$, a 1-form with values
in the Lie algebra ${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$ (or, more generally,
${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}^\NC$), where $A^{10}=t^a\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A^a_z dz$ and $A^{01}=t^a\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
A^a_{\bar z}d\bar z\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$
are, respectively, a 1,0- and a 0,1-form (the chiral components
of the gauge field). In most what follows, we shall treat the gauge
field as external, i.e. non-dynamical.
Nevertheless, the coupling will allow to test the variation
of the quantum system under the changes of the gauge field
background and, finally, will facilitate the exact solution
of the model. For a surface without boundary, we define
\begin{eqnarray}} \newcommand{\de}{\bar\partial
S(g,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A)\ =\ S(g)\,+\,
{_{ik}\over^{2\pi}}\int_{_\Sigma}
\hspace{-0.08cm}\tr\,[A^{10}(g^{-1}\de g)\,+\,
(g\partial} \newcommand{\ee}{{\rm e} g^{-1})\,A^{01}
\,+\,g\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A^{10} g^{-1}A^{01}]\,.\hspace{0.7cm}
\label{tog}
\qqq
Under the local gauge transformations $h:\Sigma
\rightarrow G$, the gauge fields transform in the standard way:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
A^{10}\,\mapsto\,{}^h\hspace{-0.09cm}A^{10}\,=\, h\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A^{10}\, h^{-1}\,
+\,h\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\partial} \newcommand{\ee}{{\rm e} h^{-1}\,,\qquad A^{01}\,\mapsto\,{}^h\hspace{-0.09cm}
A^{01}\,=\, h\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A^{01} h^{-1}\,+\,h\de h^{-1}\,.
\nonumber} \newcommand{\no}{\noindent
\qqq
The reaction of the action to the chiral changes of the gauge
is encoded in the identity
\begin{eqnarray}} \newcommand{\de}{\bar\partial
S(g,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A)\ =\ S(h_1g\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} h_2^{-1},\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{}^{h_2}
\hspace{-0.09cm}A^{10}+
{}^{h_1}\hspace{-0.1cm}A^{01})\,+\,S(h_2,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A^{10})\,
+\,S(h_1^{-1},\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A^{01})
\label{trprop}
\qqq
which follows in a direct manner from the Polyakov-Wiegmann
formula (\ref{pwWZW}). For the later convenience,
we have chosen a modified way of coupling to the gauge field,
as compared to the more standard way with the addition
of the term $-A^{10}A^{01}$ in the brackets on the right hand
side of Eq.\,\,(\ref{tog}). The latter way would render
the action invariant with respect to the diagonal
(i.e.\,\,non-chiral) gauge transformations with $h_1=h_2$.
\vskip 0.3cm
For surfaces with boundary, we define $\ee^{-S(g,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A)}$
by the same prescription but on the level of the amplitudes
with values in the product of line bundles. The definition
(\ref{prKM}) of the product implies then the transformation
rule:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\ee^{S(g,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A)}\ =\
\ee^{-S(h_1^{-1},\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A^{01})}\cdot\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\ee^{-S(h_1g\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} h_2^{-1},
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{}^{^{h_2}}\hspace{-0.11cm}A^{10}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}+\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{}^{^{h_1}}
\hspace{-0.11cm}A^{01})}\cdot\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
\ee^{-S(h_2,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A^{10})}
\label{trpropb}
\qqq
which extends the property (\ref{trprop}) to the case with
boundaries.
\vskip 0.4cm
\nsection{Quantization of the WZW model}
\subsection{Quantum amplitudes}
The Feynman quantization prescription instructs us that
in the quantum WZW model we should sum the amplitudes
of different classical configurations.
This leads to formal functional integrals such as, for example,
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\CA_{_\Sigma}\ =\ \int\ee^{-S_\Sigma(g)}\,\, Dg\,,
\nonumber} \newcommand{\no}{\noindent
\qqq
where one integrates over the maps $g:\Sigma\rightarrow G$
and $Dg$ stands for the local product $\prod_{_\xi}\hspace{-0.08cm}
dg(\xi)$ of the Haar measures. If $\Sigma$ is closed, then the above
integral should take a numerical value ${\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_\Sigma}$ called
the {\bf partition function}
(because of the statistical physics analogy). For $\Sigma$ with
boundary, it should define, instead, a Hilbert space state.
Let $\Gamma({\cal L}} \newcommand{\CM}{{\cal M})$ denote the space of sections of the line bundle
${\cal L}} \newcommand{\CM}{{\cal M}$ over the loop
group $LG$. $\,\Gamma({\cal L}} \newcommand{\CM}{{\cal M})$ plays the role of the space of states
of the quantized theory. If $\Sigma$ has a boundary, we should
fix in the functional integration the boundary
values $\un{g}=(g_n)$ of fields $g:\Sigma\rightarrow G\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\CA_{_\Sigma}(\un{g})\ =\ \int\limits_{g|_{_{\partial} \newcommand{\ee}{{\rm e} D_{_n}}}=\,g_n}
\ee^{-S_\Sigma(g)}\,\, Dg\,.
\nonumber} \newcommand{\no}{\noindent
\qqq
The result, in its dependence on $\un{g}$,
should give an element of the tensor product $\,\mathop{\otimes}
\limits_n\Gamma({\cal L}} \newcommand{\CM}{{\cal M})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ of the state spaces: the {\bf quantum amplitude}
corresponding to the surface $\Sigma$. More generally,
we shall consider the quantum amplitudes in the presence
of external gauge field:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\CA_{_\Sigma}(\un{g};\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A)\,\equiv\,
\ =\ \int\limits_{g|_{_{\partial} \newcommand{\ee}{{\rm e} D_{_n}}}=\,g_n}
\ee^{-S_\Sigma(g,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A)}\,\, Dg
\label{FfiA}
\qqq
again with the values in $\,\mathop{\otimes}
\limits_n\Gamma({\cal L}} \newcommand{\CM}{{\cal M})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$. We would like to give a rigorous
meaning to such objects. In general, the functional
integrals require complicated renormalization procedures
which, besides, work only in some cases (of the so called
renormalizable theories) and even then, in most instances,
have been implemented only on the level of formal
perturbation series. The WZW models are perturbatively
renormalizable. In this case, however, one may follow
a shortcut by exploiting formal symmetry properties
of the functional integrals and showing that they
fix uniquely the quantum amplitudes. This will be
the line of thought adopted below, although
we shall only describe the essential points of the
argument and make detours to introduce other important
notions.
\vskip 0.3cm
Let us start by discussing the formal structure
of the space of states $\Gamma({\cal L}} \newcommand{\CM}{{\cal M})$. The scalar product
and the bilinear form
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{\rm\bf(}\psi,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\psi'{\rm\bf)}\ =\ \int_{_{LG}}(\psi(g),
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\psi'(g))\,\, Dg\,,
\qquad{\bf\langle}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\psi,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\psi'\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\bf\rangle}
=\ \int_{_{LG}}\langle\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\psi(g),\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
\psi'(\check{g})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\rangle\,\,Dg\,,
\label{fbf}
\qqq
which employ the hermitian structure and the duality (\ref{glue})
on the line bundle ${\cal L}} \newcommand{\CM}{{\cal M}$, should turn $\Gamma({\cal L}} \newcommand{\CM}{{\cal M})$
into a Hilbert space ${\cal H}} \newcommand{\CI}{{\cal I}$
and should allow the identification of ${\cal H}} \newcommand{\CI}{{\cal I}$ with its dual.
The space $\Gamma({\cal L}} \newcommand{\CM}{{\cal M})$ carries also two commuting actions
of the group $\,\hat{G}\,$: $\,
\psi\,\mapsto\,{}^{\hat h}\hspace{-0.02cm}\psi\,$ and
$\,\psi\,\mapsto\,\psi^{\hspace{0.01cm}\hat h}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$.
They are defined by:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{}^{^{\hat h}}\hspace{-0.06cm}\psi(g)\
=\ \hat h\cdot\psi(h^{-1}g)\,
\qquad
\psi^{^{\hat h}}(g)\ =\ \psi(gh)\cdot{\hat h}^{-1}\,,
\qqq
where $g$ and $h$ are elements of the loup group $LG$
and $h$ is the projection of $\,\hat h\in\hat{G}$.
Formally, these actions preserve the scalar product and
the bilinear form in $\Gamma({\cal L}} \newcommand{\CM}{{\cal M})$.
On infinitesimal level, they give rise to two commuting
actions of the current algebra $\hat{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$
in $\Gamma({\cal L}} \newcommand{\CM}{{\cal M})$. We shall
denote by $J^a_n$ and $\widetilde} \def\hat{\widehat J^a_n$ the operators
in $\Gamma({\cal L}} \newcommand{\CM}{{\cal M})$ corresponding to the left and right
action of the generators $t^a_n$ of $\hat{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$.
The central generator $K$ acts
in $\Gamma({\cal L}} \newcommand{\CM}{{\cal M})$ as multiplication by $k$. Of course,
$J^a_n$ and $\widetilde} \def\hat{\widehat J^a_n$ satisfy the commutation
relation (\ref{KM}).
\vskip 0.3cm
As stressed by Segal in \cite{Segal}, there are two
important properties of the quantum amplitudes
$\CA_{_\Sigma}$ which are crucial for their rigorous
construction. The first one, is the gluing property
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\CA_{_{\widetilde} \def\hat{\widehat \Sigma}}(g_n,g_{n'})\ =\
\int \langle \CA_{_\Sigma}(g_n,g_{n_0})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},
\,\CA_{_{\Sigma'}}(\check g_{n_0},
g_{n'})\rangle\ Dg_{n_0}
\label{glueq}
\qqq
which states that for a surface $\widetilde} \def\hat{\widehat\Sigma$ glued along
boundary components of two pieces $\Sigma$ and $\Sigma'$,
as in Fig.\,\,7, the functional integral may be computed
iteratively, by first keeping the values of $g$ fixed
on the gluing circle and integrating over them only
after the integration over the fields on $\Sigma$
and $\Sigma'$, see Eq.\,\,(\ref{glue}).
Using the bilinear form on $\Gamma({\cal L}} \newcommand{\CM}{{\cal M})$
applied to the glued channel, we may write this relation
as the identity
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\CA_{_{\widetilde} \def\hat{\widehat \Sigma}}\ =\
{\bf\langle}\,\CA_{_\Sigma}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},\,\CA_{_{\Sigma'}}\,{\bf\rangle}\,.
\label{glueA}
\qqq
It is often more convenient to view the quantum
amplitudes $\CA_{_\Sigma}$ as operators from the tensor product
of some of the boundary spaces ${\cal H}} \newcommand{\CI}{{\cal I}$ into the others.
This is always possible because of the linear isomorphism between
${\cal H}} \newcommand{\CI}{{\cal I}$ and its dual. Then Eq.\,\,(\ref{glueA}) may be simply
rewritten with the use of the product of operators:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\CA_{_{\widetilde} \def\hat{\widehat \Sigma}}\ =\
\CA_{_\Sigma}\,\CA_{_{\Sigma'}}\,.
\label{glueB}
\qqq
One may also glue two boundary components in a single
connected surface $\Sigma$. The amplitude for the resulting
surface $\widetilde} \def\hat{\widehat\Sigma$ is then obtained from that
of $\Sigma$ by pairing the two corresponding factors
in the product of the Hilbert spaces or, in the operator
interpretation, by the partial trace applied to the
glued channel:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\CA_{_{\widetilde} \def\hat{\widehat \Sigma}}\ =\ \tr_{_{\cal H}} \newcommand{\CI}{{\cal I}}\,\CA_{_\Sigma}\,.
\label{glueC}
\qqq
In fact, as pointed out in \cite{Segal}, the last relation
encompasses also the previous
one if one introduces the amplitudes for
the disconnected Riemann surfaces defining them
as the tensor product of the amplitudes of the components.
Clearly, similar gluing relation should hold
for the amplitudes in external gauge field.
\vskip 0.3cm
The second important property of the quantum amplitudes
follows formally from the transformation
property (\ref{trpropb}) of the classical amplitudes
under the chiral gauge transformations
$h_{1,2}:\Sigma\rightarrow G$. It reads:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{}^{^{\hat h_1}}\hspace{-0.11cm}
\CA_{_\Sigma}(A)^{^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\hat h_2}}\ =\ \CA_{_\Sigma}(
{}^{h_2}\hspace{-0.11cm}A^{10}+\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{}^{h_1}
\hspace{-0.11cm}A^{01})
\label{trprq}
\qqq
for $\,\hat h_1^{-1}=\ee^{-S_{_\Sigma}(h_1^{-1},\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A^{01})}\,$
and $\,\hat h_2=\ee^{-S_{_\Sigma}(h_2,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A^{10})}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$.
\,The identity (\ref{trprq}) expresses the covariance
of the quantum amplitudes under the chiral gauge
transformations. It is at the basis of the rich
symmetry structure of the quantized WZW theory.
\vskip 0.4cm
\subsection{The spectrum}
To give a rigorous construction of the Hilbert space ${\cal H}} \newcommand{\CI}{{\cal I}$
of the WZW model, whose vectors represent quantum states
of a closed string moving on the group manifold,
one may resort to the representation theory of the current
algebras. The algebra $\hat{\bf g}$ possesses a distinguished
family of irreducible unitary representations labeled by pairs
$\hat R=(k,R)$, where $k$, a non-negative integer called
the level, is the value taken in the representation
by the central generator $K$ of $\hat{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$ and where $R$ is
an irreducible representation of $G$ (and of ${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$).
The irreducible unitary representations $\hat R$
act in spaces $V_{_{\hat R}}$ possessing a (unique) subspace
$V_{_R}\subset V_{_{\hat R}}$ annihilated by all
the generators $t^a_n$ with $n>0$ and carrying the irreducible
representation $R$ of the subalgebra ${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}\subset\hat{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$
generated by $t^a_0$. They are characterized by this property.
Not all irreducible representations $R$ of $G$ appear
for the fixed level $k$ but only the ones corresponding to the
the so called integrable HW's which satisfy the condition
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\tr\,(\phi^\vee\lambda_{_R})\,\leq\, k\,,
\label{iHW}
\qqq
where $\phi=\phi^\vee$ is the highest root of ${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$, i.e.
such a root that $\phi+\alpha$ is not a root for any positive
root $\alpha$. Given $k$, there is only a finite
number of integrable HW's.
For $\hat{su(2)}$, the integrable HW's correspond
to spins $j=0,{1\over 2},1,\dots,{k\over2}$.
If $\lambda_{_R}$ satisfies the condition (\ref{iHW})
then so does $\lambda_{_{\overline{R}}}$ and the space
$V_{_{\hat{\overline{R}}}}$ is canonically isomorphic to
$\overline{V_{_{\hat{R}}}}$. The scalar product
on $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} V_{_{\hat{R}}}$ may then be viewed as a bilinear pairing
between $V_{_{\hat{R}}}$ and $V_{_{\hat{\overline{R}}}}$.
\vskip 0.3cm
The rigorous definition of the Hilbert space of states ${\cal H}} \newcommand{\CI}{{\cal I}$
for the WZW model of level $k$ makes the two notions of the level
coincide:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{\cal H}} \newcommand{\CI}{{\cal I}\ \ =\ \mathop{\oplus}\limits_{\hat R\ {\rm of\ level}\ k}
\left(V_{_{\hat R}}\otimes V_{_{\hat{\overline{R}}}}\right)^{-}\,,
\label{HSrd}
\qqq
where the symbol $\,(\dots)^-$ stands for the Hilbert space
completion. This is the loop group analogue of the decomposition
(\ref{decomp}) of $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} L^2(G,dg)$.
The operators $J^a_n$ and $\widetilde} \def\hat{\widehat J^a_n$ representing the action
of the generators $t^a_n$ in, respectively, $V_{_{\hat R}}$
and $V_{_{\hat{\overline{R}}}}$ satisfy the unitarity conditions
$\,{J^a_n}^\dagger=J^a_{-n}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\widetilde} \def\hat{\widehat J^a_n{}^\dagger=
\widetilde} \def\hat{\widehat J^a_{-n}$. It is not difficult to motivate
the above choice of the Hilbert space. \,One may, indeed,
realize the space (\ref{HSrd}) as a space of sections
of ${\cal L}} \newcommand{\CM}{{\cal M}$. Formally, this may be done by the assignment
(compare to the relation (\ref{hwq}))
\begin{eqnarray}} \newcommand{\de}{\bar\partial
V_{_{\hat R}}\otimes V_{_{\hat{\overline{R}}}}\supset
V_{_R}\otimes V_{_{\overline{R}}}\,\ni\,e^i_{_R}\otimes\overline{e^j_{_R}}
\ \ \ \mapsto\ \ \ \zeta
\int\overline{g_{_R}^{^{ij}}(0)}
\,\,\ee^{-S_{_D}(g)}\,\, Dg\,,
\label{HWq}
\qqq
where the normalization constant $\zeta$ will be fixed later.
The functional integral on the right hand side,
as a function of $g\vert_{_D}=h$ is the corresponding section
of ${\cal L}} \newcommand{\CM}{{\cal M}$. One may argue that the above
integral is given, up to normalization, by its semi-classical
value,
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\overline{(g_{cl})_{_R}^{^{ij}}(0)}\,\,\ee^{-S_{_D}(g_{cl})}
\label{HWcl}
\qqq
where $g_{cl}:D\rightarrow G^\NC$ is the solution of the
classical equations (\ref{cleq}) with the boundary condition
$g_{cl}\vert_{_{\partial} \newcommand{\ee}{{\rm e} D}}=h$. As shown in \cite{7}, the expression
(\ref{HWcl}) defines a non-singular section of ${\cal L}} \newcommand{\CM}{{\cal M}$
only if the HW of $R$ is integrable at level $k$ (recall
that the action $S(g_{cl})$ is proportional to $k$).
One obtains then a rigorous embedding of the space
$\,\mathop{\oplus}_{\hat R}V_R\otimes V_{\overline{R}}\,$
into $\Gamma({\cal L}} \newcommand{\CM}{{\cal M})$, and, identifying the actions of the
current algebra $\hat{{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}}$ in $\Gamma({\cal L}} \newcommand{\CM}{{\cal M})$ and
in $\,\mathop{\oplus}_{\hat R}V_{\hat R}
\otimes V_{\hat{\overline{R}}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$, \,also of the latter space.
The formal scalar product and the formal bilinear form
(\ref{fbf}) on $\Gamma({\cal L}} \newcommand{\CM}{{\cal M})$ correspond to the scalar product
and the bilinear form on ${\cal H}} \newcommand{\CI}{{\cal I}$ induced by the scalar
product on the representation spaces $V_{_{\hat{R}}}$
and the bilinear pairing between $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} V_{_{\hat{R}}}$
and $V_{_{\hat{\overline{R}}}}$ (the latter induces the pairing
between the $\hat R$ and $\hat{\overline{R}}$ summands in ${\cal H}} \newcommand{\CI}{{\cal I}$).
\vskip 0.3cm
The action of the pair of the current algebras in ${\cal H}} \newcommand{\CI}{{\cal I}$
leads to the (projective) action in ${\cal H}} \newcommand{\CI}{{\cal I}$ of the algebra
of conformal symmetries. Let us discuss how this occurs.
The spaces $V_{\hat R}$ of the irreducible unitary representations
$\hat R$ of $\hat{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$ appear to carry also the unitary representations
of the {\bf Virasoro algebra} $Vir$, the central extension of the
algebra of the vector fields $Vect(S^1)$ on the circle,
\begin{eqnarray}} \newcommand{\de}{\bar\partial
0\rightarrow\NR\rightarrow Vir\rightarrow Vect(S^1)\rightarrow
0\,.
\label{vir}
\qqq
The complex generators $\ell_n$ of $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} Vir$ corresponding to
the vector fields $\,i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} in\varphi}{\partial} \newcommand{\ee}{{\rm e}_\varphi}\,$ satisfy
the commutation relations
\begin{eqnarray}} \newcommand{\de}{\bar\partial
[\ell_n,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\ell_m]\ =\ (n-m)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\ell_{n+m}\,+\,{_1\over^{12}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
C\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}(n^3-n)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\delta_{n+m,0}\,,
\label{Vir}
\qqq
where $C$ is the central generator, the image of $1$ under
the second arrow in the exact sequence (\ref{vir}).
The action of the generators $\ell_n$ in the spaces of the
representations $\hat R$ of $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\hat{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ gives rise to the
set of operators $L_n$ and $\widetilde} \def\hat{\widehat L_n$ acting in the space
of states $\,\oplus_{_{\hat R}}V_{_{\hat R}}\otimes
V_{_{\hat{\overline{R}}}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}.$ They implement a projective action
of $\,Vect(S^1)\oplus Vect(S^1)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$, \,the Lie algebra
of Minkowskian conformal transformations.
$\,Vect(S^1)\oplus Vect(S^1)\,$ is, indeed, the
Lie algebra of the infinitesimal transformations preserving
the conformal class of the Minkowski metric $dx^2-dt^2=dx^+dx^-$,
where $\,x^\pm=x\pm t\,$ are the light-cone coordinates
on the cylinder with periodic space-coordinate $x$.
\vskip 0.3cm
Explicitly, the operators $L_n$'s and $\widetilde} \def\hat{\widehat L_n$'s
are given in terms of the operators $J^a_n$
and $\widetilde} \def\hat{\widehat J^a_n$, generating the actions of $\hat{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$,
by the so called Sugawara construction:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
L_n\ =\ {_1\over^{k+h^\vee}}\sum\limits_{m=-\infty}^\infty
J^a_{n-m}J^a_{m}\quad\ {\rm for}\ \ n\not=0\,,\qquad
L_0\ =\ {_2\over^{k+h^\vee}}\sum\limits_{m=0}^\infty
J^a_{-n}J^a_n
\label{sug}
\qqq
and similarly for $\widetilde} \def\hat{\widehat L_n$. Above, $h^\vee$ (the {\bf dual Coxeter
number}) stands for the value of the quadratic Casimir in the adjoint
representation of ${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$ and is equal to $N$ for the $SU(N)$
group. The operators $L_n$ and $\widetilde} \def\hat{\widehat L_n$
satisfy the relations (\ref{Vir}) with $C$ acting as the multiplication
by $c={k\,{\rm dim}(G)\over{k+h^\vee}}$, the value
of the {\bf Virasoro central charge} of the WZW theory. Besides,
\begin{eqnarray}} \newcommand{\de}{\bar\partial
[L_n,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} J^a_m]\ =\ -m J^a_{n+m}
\nonumber} \newcommand{\no}{\noindent
\qqq
and similarly for $[\widetilde} \def\hat{\widehat L_n,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\widetilde} \def\hat{\widehat J^a_m]$. The operators
$L_n$ (and $\widetilde} \def\hat{\widehat L_n$) satisfy the unitarity conditions
$L_n^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\dagger}=L_{-n}$. In particular $L_0$ is a self-adjoint
operator, bounded below on $V_{_{\hat R}}$ by
the {\bf conformal dimensions} $\Delta_{_R}={c_{_R}\over k+h^\vee}$,
the eigenvalue of $L_0$ on the subspace $V_{_R}\subset V_{_{\hat R}}$.
The latter subspace is annihilated by all $L_n$ with $n>0$.
The Hamiltonian of the
WZW theory is $\,H=L_0+\widetilde} \def\hat{\widehat L_0-{c\over 12}\,$ whereas
$L_0-\widetilde} \def\hat{\widehat L_0$ defines the momentum operator $P$. The
tensor product of the HW vectors in the subspace
$\,V_{_{\hat 1}}\otimes\overline{V}_{_{\hat 1}}\subset{\cal H}} \newcommand{\CI}{{\cal I}\,$
corresponding to the trivial representation $R=1$
gives the {\bf vacuum state} $\Omega$ of the theory annihilated
by $L_0$ and $\widetilde} \def\hat{\widehat L_0$.
\vskip 0.3cm
A certain role in what follows will be played by the
characters of the representations $V_{_{\hat R}}$
defined as traces of loop group operators
acting in $V_{_{\hat R}}$. To avoid domain problems,
one often considers only the endomorphisms
$g_{_{\hat R}}$ of $V_{_{\hat R}}$ representing the action
of the elements $g\in G$ (or in $G^\NC$)
and obtained by the integration of the action of the
generators $t^a_0$. One then defines
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\chi_{_{\hat R}}(\tau,g)\ =\ \tr_{_{V_{_{\hat R}}}}
\ee^{\m2\pi i\tau\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}(L_0-{c\over 24})}\, g_{_{\hat R}}\,,
\label{hatch}
\qqq
where $\tau$ is a complex number in the upper half plane:
${\rm Im}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\tau>0$. The presence of the factor $\,\ee^{\m2\pi
i\tau\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} L_0}\,$ renders the trace finite. The characters
$\chi_{_{\hat R}}(\tau,g)$ are class functions of $g$
and may be explicitly computed. Their decomposition
into characters of $G$ encodes the multiplicities
of the eigenvalues of the Virasoro generator
$L_0$ in the subspaces of $V_{_{\hat R}}$ transforming
according to a given representation of $G$.
\vskip 0.3cm
The central charge $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} c\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ entering the commutation
relations of the Virasoro generators is an important
characteristic of a conformal field theory.
It appears also in the rigorous version of the quantum amplitudes
$\CA_{_\Sigma}$. It enters into them in a somewhat subtle way,
measuring their change under the local rescalings $\eta\mapsto
\ee^{2\sigma}\eta$ of the metric of $\Sigma$ (recall that
the amplitudes of the classical configurations $\ee^{-S(g)}$
were invariant under such rescalings). Under the change
$\eta\mapsto\ee^{2\sigma}\eta$ with $\sigma$ vanishing
around the boundary,
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\CA_{_\Sigma}\ \ \mapsto\ \ \ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{c\over 12\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\pi}\int_{_\Sigma}
\hspace{-0.06cm}[{_1\over^2}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\partial} \newcommand{\ee}{{\rm e}_\alpha\sigma\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\partial} \newcommand{\ee}{{\rm e}_\beta\sigma\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
\eta^{\alpha\beta}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}+\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\sigma\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} r\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}+\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\mu(\ee^{2\sigma}-1)]
\sqrt{\eta}}\ \CA_{_\Sigma}
\label{scin}
\qqq
due to renormalization effects, with $\mu$ depending on
the renormalization prescription. We shall use the prescription
corresponding to $\mu=0$. The same transformation rule is obeyed
by the amplitudes $\CA_{_\Sigma}(A)$ in external gauge field.
Hence, the quantum amplitudes are only projectively
invariant under the conformal
rescalings of the metric $\eta$. This is an example
of the standard effect leading to projective actions
of symmetries in quantum theory. Due to this effect, some
care will have to be taken when making sense out of
formal properties of the quantum amplitudes, like
the gluing property (\ref{glueq}). We shall always assume
that the metric $\eta$ of $\Sigma$, which, together
with the orientation of $\Sigma$, defines its complex structure,
is of the special form around the boundary.
Namely, that, in terms of the complex coordinate of
the unit discs $D_{n}$ holomorphically embedded into
the surface $\Sigma'$ without boundary such that
$\,\Sigma=\Sigma'\setminus(\mathop{\cup}\limits_n
D_{n}^{^{^{{\hspace{-0.21cm}}o}}})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$, it is equal to
the cylindrical metric $\vert z\vert^{-2}\vert dz\vert^2$.
Upon gluing of surfaces along boundary components, such metrics
will automatically give smooth metrics on the resulting
surfaces.
\vskip 0.3cm
In particular, the metrics on the unit discs $D$
will have the form $\ee^{2\sigma}\vert dz\vert^2$
with $\sigma=-\ln{\vert z\vert}$ around the boundary
of $D$. Unless otherwise stated, we shall also assume
that $\sigma=0$ around the center of $D$.
Consider the Riemann sphere $\NC P^1=\NC\cup\{\infty\}$
composed from the two copies of the disc $D$ glued along
the boundary. The choice
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\zeta\ =\ {\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{\NC P^1}^{-{_1\over^2}}
\label{ncst}
\qqq
for the normalizing constant will make precise
the assignment (\ref{HWq}). This choice guarantees
that the change of $\zeta$ under the rescalings of the metric
on $D$ will cancel the change of the functional integral
$\,\int\overline{g^{ij}_{_R}}\,\ee^{-S(g)}\,Dg\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$.
\vskip 0.4cm
\subsection{Correlation functions}
The formalism of Green functions encoding the action of field
operators constitutes a traditional tool in quantum field theory.
In the Minkowski space, the Green functions allow
to express easily the scattering matrix elements
(at least for the massive theories, via the LSZ formalism)
whereas in the Euclidean space they
coincide with correlation functions of continuum
statistical models, providing a bridge between quantum
field theory and statistical mechanics. In the context
of CFT, the correlation functions defined on a general
Riemann surface $\Sigma$ without boundary constitute
a somewhat easier objects to deal with than
the quantum amplitudes $\CA_{_\Sigma}$ for surfaces with
boundary. Besides, even considered on the simplest Riemann
surface, the Riemann sphere $\NC P^1$,
they already contain the full information about the model.
Formally, the correlation functions of the WZW model
are given by the functional integrals
\begin{eqnarray}} \newcommand{\de}{\bar\partial
<\prod\limits_n g^{i_nj_n}_{_{R_n}}(\xi_n)>_{_{\hspace{-0.08cm}\Sigma}}
\hspace{-0.1cm}(A)\ =\ {\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_\Sigma}(A)^{-1}
\int\prod\limits_n{g_{_{R_n}}^{i_nj_n}}\ \ee^{-S_{_\Sigma}
(g,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A)}\,\,Dg\,,
\label{cfu}
\qqq
where $\xi_n$ are disjoint points in a Riemann surface $\Sigma$
without boundary. For the Riemann surface without boundary
$\Sigma'$ obtained by gluing unit discs $D_{n}$ to a surface $\Sigma$
with boundary\footnote{the metric $\eta$ on $\Sigma'$
is assumed to come from metrics on $\Sigma$ and on the discs
$D_{n}$ of the type described above}, \hspace{0.025cm}} \newcommand{\ch}{{\rm ch} see Fig.\,\,6,
and for the points $\xi_n$ placed at the centers of the discs
$D_{n}$, the correlation functions without the gauge field
may be expressed, with the use of the assignment (\ref{HWq})
and of the gluing property (\ref{glueA}), by the scalar products
of the quantum amplitudes $\CA_{_\Sigma}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ with special vectors
in the Hilbert space of states:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
<\prod\limits_n g^{i_nj_n}_{_{R_n}}(\xi_n)>_{_{\hspace{-0.08cm}\Sigma'}}
\ =\ {\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_\Sigma}^{-1}\,\,{\bf(}\mathop{\otimes}
\limits_{n}(e^{i_n}_{_{R_n}}\otimes\overline{e^{i_n}_{_{R_n}}})
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\CA_{_\Sigma}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\bf)}\,.
\nonumber} \newcommand{\no}{\noindent
\qqq
The normalization factor is given by the partition
function of the surface with boundary $\Sigma$ defined by
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{\Sigma}}\ =\ {\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{\Sigma'}}\prod_n\zeta_n
\qqq
with $\zeta_n$ as in Eq.\,\,(\ref{ncst}).
The combination of the partition functions
on the right hand side does not change under local rescalings
of the metric inside the discs $D_{n}$.
\vskip 0.3cm
On the level of correlation functions,
the symmetry properties of the theory are encoded
in the so called Ward identities. For example,
the behavior (\ref{trprop}) of the action
under the chiral gauge transformations with
for $h_1=h$ and $h_2=1$ implies formally that
\begin{eqnarray}} \newcommand{\de}{\bar\partial
&&{\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{\Sigma}}(A)\,<\mathop{\otimes}\limits_n g_{_{R_n}}
(\xi_n)>_{_{\hspace{-0.08cm}\Sigma}}\hspace{-0.1cm}(A)\cr\cr
&&\hspace{0.6cm}=\ \ee^{-S(h^{-1},\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A^{01})}\
\mathop{\otimes}\limits_n h^{-1}_{_{R_n}}(\xi_n)\
{\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{\Sigma}}(A^{10}+{}^{h}\hspace{-0.09cm}A^{01})
\,<\mathop{\otimes}\limits_n g_{_{R_n}}(\xi_n)
>_{_{\hspace{-0.08cm}\Sigma}}\hspace{-0.1cm}
(A^{10}+{}^{h}\hspace{-0.09cm}A^{01})\,,\hspace{1cm}
\label{WIcfl}
\qqq
where we view $\,\otimes g_{_{R_n}}\,$
as taking value in $\,\mathop{\otimes}\limits_nEnd(V_{_{R_n}})$
and collecting all the matrix elements $\,\prod g^{^{i_nj_n}}
_{_{R_n}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$. \,Similarly, for $h_1=1$ and $h_2=h$, we obtain
the mirror relation:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
&&{\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{\Sigma}}(A)\,<\mathop{\otimes}\limits_n g_{_{R_n}}
(\xi_n)>_{_{\hspace{-0.08cm}\Sigma}}\hspace{-0.1cm}(A)\cr\cr
&&\hspace{0.6cm}=\ \ee^{-S(h,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A^{10})}
\ {\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{\Sigma}}({}^{h}\hspace{-0.09cm}A^{10}+A^{01})\,
<\mathop{\otimes}\limits_n g_{_{R_n}}(\xi_n)
>_{_{\hspace{-0.08cm}\Sigma}}\hspace{-0.1cm}
({}^{h}\hspace{-0.09cm}A^{10}+A^{01})
\ \mathop{\otimes}\limits_n h_{_{R_n}}(\xi_n)\,.\hspace{0.8cm}
\label{WIcfr}
\qqq
These are the Ward identities expressing the symmetry
of the correlation functions under the chiral gauge
transformations.
\vskip 0.3cm
It is useful and customary to introduce more general
correlation functions with insertions of {\bf currents} testing
the reaction of the functions (\ref{cfu})
to infinitesimal changes of the gauge fields.
On the surface $\Sigma'$ they are defined by
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{\Sigma'}}<J^a(z_m)\prod\limits_n g^{^{i_nj_n}}_{_{R_n}}
(\xi_n)>_{_{\hspace{-0.08cm}\Sigma'}}
\,=\,-\pi\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{{\delta}\over{\delta A^a_{\bar z}(z_m)}}
{\bigg\vert}_{_{A=0}}{\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{\Sigma'}}\hspace{-0.1cm}(A)
<\prod\limits_n g^{i_nj_n}_{_{R_n}}
(\xi_n)>_{_{\hspace{-0.08cm}\Sigma'}}\hspace{-0.1cm}(A)\,,
\hspace{0.6cm}
\qqq
where $z_m$ is the complex coordinate of the disc $D_m$, or by
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{\Sigma'}}<\widetilde} \def\hat{\widehat J^a(\bar z_m)\prod\limits_n
g^{^{i_nj_n}}_{_{R_n}}(\xi_n)>_{_{\hspace{-0.08cm}\Sigma'}}
\,=\,-\pi\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{{\delta}\over{\delta A^a_{z}(z_m)}}
{\bigg\vert}_{_{A=0}}{\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{\Sigma'}}(A)
<\prod\limits_n g^{i_nj_n}_{_{R_n}}(\xi_n)
>_{_{\hspace{-0.08cm}\Sigma'}}(A)\,.\hspace{0.6cm}
\qqq
Its is not very difficult to show, expanding the
Ward identities (\ref{WIcfl}) and (\ref{WIcfr})
to the first order in $h$ around $1$,
that the insertions of $J^a(z)$ ($\widetilde} \def\hat{\widehat J^a(\bar z)$)
are analytic (anti-analytic) in $z\not=0$ but that
they have simple poles at $z=0$, the location point
of one of the insertions $g_{_{R}}(\xi)$,
with the behavior
\begin{eqnarray}} \newcommand{\de}{\bar\partial
J^a(z)\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g_{_{R}}(\xi)\ =\ -\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{1\over{z}}\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} t^a_{_R}\,
g_{_{R}}(\xi)\ +\ \dots\,,\qquad
\widetilde} \def\hat{\widehat J^a(\bar z)\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g_{_{R}}(\xi)\ =\ {1\over{\bar z}}\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
g_{_{R}}(\xi)\, t^a_{_R}\ +\ \cdots\,,\hspace{0.8cm}
\label{fope}
\qqq
where the dots denote non-singular terms. The latter are
related to the action of the current algebra generators
in the space of states by the following relations involving
the contour integrals\footnote{oriented counter-clockwise}:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{_1\over^{2\pi i}}\hspace{-0.2cm}\int\limits_{\vert z_m\vert=\rho}
\hspace{-0.3cm}
<J^a(z_m)\prod\limits_n g^{i_nj_n}_{_{R_n}}(\xi_n)
>_{_{\hspace{-0.08cm}\Sigma'}}
\, z_m^p\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} dz_m&=&
-\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_\Sigma}^{-1}\,\,{\bf(}\mathop{\otimes}
\limits_{n}({J^a_{p}}^{^{\hspace{-0.05cm}\#}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} e^{i_n}_{_{R_m}}
\otimes
\overline{e^{i_n}_{_{R_n}}})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\CA_{_\Sigma}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\bf)}\,,
\label{cfl}\hspace{1.3cm}\\
-\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{_1\over^{2\pi i}}\hspace{-0.2cm}\int\limits_{\vert z_m\vert=\rho}
\hspace{-0.3cm}
<\widetilde} \def\hat{\widehat J^a(z_m)\prod\limits_n g^{i_nj_n}_{_{R_n}}(\xi_n)
>_{_{\hspace{-0.08cm}\Sigma'}}
\,\bar z_m^p\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} d\bar z_m&=&
-\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_\Sigma}^{-1}\,\,{\bf(}\mathop{\otimes}
\limits_{n}(e^{i_n}_{_{R_n}}\otimes
\widetilde} \def\hat{\widehat J^a_{p}{}^{^{\hspace{-0.05cm}\#}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
\overline{e^{i_n}_{_{R_n}}})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
\CA_{_\Sigma}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\bf)}\hspace{1.3cm}
\label{cfr}
\qqq
with $\rho<1$ and the superscript $\#$ indicating that the operator
appears only for $n=m$. Eqs.\,\,(\ref{fope}) are examples
of the {\bf operator product expansions}, in this case, the ones
stating that $g_{_{R}}(\xi)$ are {\bf primary fields} of the
current algebra, in the CFT jargon.
\vskip 0.3cm
Multiple current insertions, see Fig.\,\,8, integrated over contours
of increasing radii lead to multiple insertions of the current algebra
generators. For example:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{_1\over^{2\pi i}}\hspace{-0.2cm}\int\limits_{\vert z_m\vert=\rho_1}
\hspace{-0.3cm}
{_1\over^{2\pi i}}\hspace{-0.2cm}\int\limits_{\vert w_m\vert=\rho_2}
\hspace{-0.3cm}
<J^a(z_m)\,J^b(w_m)
\prod\limits_n g^{i_nj_n}_{_{R_n}}(\xi_n)
>_{_{\hspace{-0.08cm}\Sigma'}}\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} z_m^{p}\,w_m^q
\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} dz_m\, dw_m\hspace{3cm}\cr
=\ {\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_\Sigma}^{-1}\,\,{\bf(}\mathop{\otimes}
\limits_{n}(J^a_{p}{}^{^{\hspace{-0.05cm}\#}}
J^b_{q}{}^{^{\hspace{-0.05cm}\#}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} e^{i_n}_{_{R_n}}
\otimes\overline{e^{i_n}_{_{R_n}}})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\CA_{_\Sigma}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\bf)}
\hspace{1cm}
\nonumber} \newcommand{\no}{\noindent
\qqq
for $\rho_1>\rho_2$ whereas for $\rho_1<\rho_2$
the order of $J^a_{p}J^b_{q}$
should be reversed. In particular,
the commutator $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}[J^a_{p},\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} J^b_{q}]\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ corresponds
to the difference of the two double contour integrals.
It follows, that general matrix elements
of the quantum amplitudes $\CA_\Sigma$ may be read
of the correlation function in the external gauge. It is then
enough to find the latter to describe the complete theory.
\leavevmode\epsffile[-80 -20 337 190]{fig8.eps}
\vskip 0.1cm
The action of the Virasoro generators may be interpreted
similarly in terms of the insertions of the {\bf energy-momentum}
tensor into the correlation functions which test their variation
under the changes of the metric on the surface:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{\Sigma'}}\,<T(z_m)\prod\limits_n g^{i_nj_n}_{_{R_n}}(\xi_n)
>_{_{\hspace{-0.08cm}\Sigma'}}
\,&=&\,4\pi\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{{\delta}\over{\delta\eta^{zz}(z_m)}}\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
{\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{\Sigma'}}\,
<\prod\limits_n g^{i_nj_n}_{_{R_n}}(\xi_n)
>_{_{\hspace{-0.08cm}\Sigma'}}\,,
\cr
{\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{\Sigma'}}\,<\widetilde} \def\hat{\widehat T(\bar z_m)\prod\limits_n
g^{i_nj_n}_{_{R_n}}(\xi_n)>_{_{\hspace{-0.08cm}\Sigma'}}
\,&=&\,4\pi\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{{\delta}\over{\delta\eta^{\bar z\bar z}(z_m)}}\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
{\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{\Sigma'}}\,<\prod\limits_n g^{i_nj_n}_{_{R_n}}(\xi_n)
>_{_{\hspace{-0.08cm}\Sigma'}}\,.
\qqq
Under the local rescaling of the metric $\eta\mapsto\ee^{2\sigma}
\eta$ with $\sigma$ vanishing around the insertion
points $\xi_n$, the correlation functions (\ref{cfu}) are invariant.
This is not any more the case for general $\sigma$ due
to (the ``wave function'') renormalization of the insertions.
For general $\sigma$, the correlation functions
pick up the product of local factors equal to
$\ee^{-2\Delta_{_{R_n}}\hspace{-0.06cm}\sigma(\xi_n)}$,
where the {\bf conformal
dimension} $\Delta_{_R}$ of the fields $g_{_R}(\xi)$
coincide with the lowest eigenvalues of the Virasoro generator
$L_0$ in the HW representations of the current algebra
discussed before. The partition functions transform under
the metric rescaling according to the rule (\ref{scin}).
The infinitesimal versions of the above transformation properties
together with the covariance of the whole scheme under
infinitesimal diffeomorphisms of the surface $\Sigma$ may be shown
to imply that the insertions of $T(z_m)$ $(\widetilde} \def\hat{\widehat T(\bar z_m)$)
are analytic (anti-analytic)\footnote{in the standard metric
$\vert dz\vert^2$ around the insertion point} in $z_m$ for $z_m\not=0$
with the singular part given by the operator product expansion
\begin{eqnarray}} \newcommand{\de}{\bar\partial
&&T(z)\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g_{_R}(\xi)\ =\ {1\over{z^2}}\,{\Delta_{_R}}\,g_{_R}(\xi)
\,+\,{1\over z}\,\partial} \newcommand{\ee}{{\rm e}_z\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g_{_R}(\xi)\ +\ \dots\,,\cr\cr
&&\widetilde} \def\hat{\widehat T(\bar z)\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g_{_R}(\xi)\ =\ {1\over{{\bar z}^2}}
\,{\Delta_{_R}}\,g_{_R}(\xi)
\,+\,{1\over{\bar z}}\,\partial} \newcommand{\ee}{{\rm e}_{\bar z}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g_{_R}(\xi)\ +\ \dots\,.
\nonumber
\qqq
The latter expansions state that $g_{_R}(\xi)$ are primary fields
of the Virasoro algebra. The insertions of the energy momentum
tensor encode the action of the Virasoro algebra generators
in the space of states:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{_1\over^{2\pi i}}\hspace{-0.2cm}\int\limits_{\vert z_m\vert=\rho}
\hspace{-0.3cm}
<T(z_m)\prod\limits_n g^{i_nj_n}_{_{R_n}}(\xi_n)
>_{_{\hspace{-0.08cm}\Sigma'}}
\, z_m^{p+1}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} dz_m&=&
{\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_\Sigma}^{-1}\,\,{\bf(}\mathop{\otimes}
\limits_{n}(L_{p}^{^{\hspace{-0.04cm}\#}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} e^{i_n}_{_{R_n}}\otimes
\overline{e^{i_n}_{_{R_n}}})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\CA_{_\Sigma}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\bf)}\,,
\hspace{1.5cm}
\label{cflV}\\
-\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{_1\over^{2\pi i}}\hspace{-0.2cm}\int\limits_{\vert z_m\vert=\rho}
\hspace{-0.3cm}
<\widetilde} \def\hat{\widehat T(\bar z_m)\prod\limits_n g^{i_nj_n}_{_{R_n}}(\xi_n)
>_{_{\hspace{-0.08cm}\Sigma'}}
\,\bar z_m^{p+1}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} d\bar z_m&=&
{\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_\Sigma}^{-1}\,\,{\bf(}\mathop{\otimes}
\limits_{n}(e^{i_n}_{_{R_n}}\otimes
\widetilde} \def\hat{\widehat L_{p}^{^{\hspace{-0.02cm}\#}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\overline{e^{i_n}_{_{R_n}}})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
\CA_{_\Sigma}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\bf)}\,.
\hspace{1.5cm}
\label{cfrV}
\qqq
\vskip 0.3cm
The Ward identities of the chiral gauge symmetry together with
the transformation properties under the local rescalings
of the metric and under diffeomorphisms of the surface,
expanded to the second order in the symmetry generators,
yield the operator product expansions
\begin{eqnarray}} \newcommand{\de}{\bar\partial
J^a(z)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} J^b(w)&=&{{k\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\delta^{ab}}\over{2(z-w)^2}}\,+\,
{{i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} f^{abc}}\over{z-w}}\, J^c(w)\,+\ \dots\,,\cr\cr
T(z)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} T(w)\,\,&=&{{c}\over{2(z-w)^4}}\,+\,
{{2}\over{(z-w)^2}}\, T(w)\,+\,
{1\over{z-w}}\, \partial} \newcommand{\ee}{{\rm e}_w T(w)\,+\ \dots\,,\cr\cr
T(z)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} J^a(w)&=&{{1}\over{(z-w)^2}}\, J^a(w)\,+\,
{1\over{z-w}}\, \partial} \newcommand{\ee}{{\rm e}_w J^a(w)\,+\ \dots\,,\cr
\label{opem}\\
\widetilde} \def\hat{\widehat J^a(\bar z)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\widetilde} \def\hat{\widehat J^b(\bar w)&=&{{k\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\delta^{ab}}
\over{2(\bar z-\bar w)^2}}\,+\,
{{i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} f^{abc}}\over{\bar z-\bar w}}\,\widetilde} \def\hat{\widehat J^c(\bar w)\,
+\ \dots\,,\cr\cr
\widetilde} \def\hat{\widehat T(\bar z)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\widetilde} \def\hat{\widehat T(w)\,\,&=&{{c}
\over{2(\bar z-\bar w)^4}}\,
+\,{{2}\over{(\bar z-\bar w)^2}}\,\widetilde} \def\hat{\widehat T(\bar w)\,+\,
{1\over{\bar z-\bar w}}\, \partial} \newcommand{\ee}{{\rm e}_{\bar w}\widetilde} \def\hat{\widehat T(\bar w)\,
+\ \dots\,,\cr\cr
\widetilde} \def\hat{\widehat T(\bar z)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\widetilde} \def\hat{\widehat J^a(\bar w)
&=&{{1}\over{(\bar z-\bar w)^2}}
\,\widetilde} \def\hat{\widehat J^a(\bar w)\,+\,{1\over{\bar z-\bar w}}\,
\partial} \newcommand{\ee}{{\rm e}_{\bar w}\widetilde} \def\hat{\widehat J^a(w)\,+\ \dots\,,
\nonumber
\qqq
where the dots denote the non-singular terms analytic
(anti-analytic) in $z$ around $w$. The above expansions
hold when inserted
into the correlation functions as above with $z$ and
$w$ corresponding to the values of the same local coordinate
for two different points and in the standard metric.
They encode through the relations (\ref{cfl}), (\ref{cfr}),
(\ref{cflV}) and (\ref{cfrV}) the commutation relations
of the current and Virasoro generators
$J_n^a,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\widetilde} \def\hat{\widehat J^a_n,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} L_n,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\widetilde} \def\hat{\widehat L_n$
obtained from the above expansions by the deformation of
the integration contours
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\int\limits_{\vert z\vert=\rho+\epsilon}dz
\int\limits_{\vert w\vert=\rho}dw
-\int\limits_{\vert z\vert=\rho-\epsilon}dz
\int\limits_{\vert w\vert=\rho}dw
\ =\ \int\limits_{\vert w\vert=\rho}dw
\int\limits_{\vert z-w\vert=\epsilon}dz\,,
\qqq
see Fig.\,\,9, and the use of the residue theorem.
\leavevmode\epsffile[-70 -20 347 155]{fig9.eps}
\noindent The operator expansion
algebra of the insertions into the correlation functions
substitutes then for the operator commutation relations
but allows to encode also more complicated algebraic
relations between the CFT operators, see the last
section. It is the basic technique of two-dimensional
CFT.
\vskip 0.3cm
As we have discussed before, the action of the Virasoro generators
in the space of states of the WZW model may be expressed
in terms of the current algebra action, see Eq.\,\,(\ref{sug}).
This relation may be translated into the language
of the insertions into the correlation functions,
giving rise to the Sugawara construction
of the energy-momentum tensor:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
T(w)\ =\ \lim\limits_{z\to w}\ {_1\over^{k+h^\vee}}\left(
J^a(z)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} J^a(w)\,-\,{_{k\,{\rm dim}(G)}\over^{2\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}(z-w)^2}}\right)
\nonumber} \newcommand{\no}{\noindent
\qqq
and similarly for $\widetilde} \def\hat{\widehat T(\bar w)$.
\vskip 0.5cm
\nsection{Chiral WZW theory and the Chern-Simons states}
As we have seen, the whole information about
the quantum amplitudes of the WZW theory resides
in the correlation functions (\ref{cfu}) in an
external gauge field. We shall look now more closely
into the gauge-field dependence of these functions.
By (formal) analytic continuation, the chiral
Ward identities (\ref{WIcfl})
and (\ref{WIcfr}) should also hold for the complexified
gauge fields $A$ with values in ${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}^\NC$
and for the complexified gauge transformations $h$ with
values in $G^\NC$. As we shall see, they give a powerful
tool for analysis of the correlation functions.
\vskip 0.3cm
Let us consider first the Ward identity (\ref{WIcfl}).
The holomorphic maps $\Psi$ on the space ${\CA}^{01}$ of
${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}^\NC$-valued 0,1-gauge fields $A^{01}$ with values in
$\,\mathop{\otimes}\limits_nV_{_{R_n}}\,$
satisfying the equation
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\Psi(A^{01})\ =\ \ee^{-S(h^{-1},\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A^{01})}\
\mathop{\otimes}\limits_n h^{-1}_{_{R_n}}(\xi_n)\
\Psi({}^{h}\hspace{-0.09cm}A^{01})
\label{CSst}
\qqq
for $h$ in the group $\CG^\NC$ of $G^\NC$-valued gauge
transformations have an interesting geometric interpretation.
On one side, they may be viewed as holomorphic sections
of a vector bundle $W$ with typical fiber
$\,\mathop{\otimes}\limits_nV_{_{R_n}}\,$
over the orbit space\footnote{this space requires
a careful definition with a special treatment of
bad orbits} ${\cal N}} \newcommand{\CO}{{\cal O}=\CA^{01}/\CG^\NC$. Mathematically, the orbit space
${\cal N}} \newcommand{\CO}{{\cal O}$ is the moduli space of the holomorphic $G^\NC$-bundles
and the mathematicians like to view $\Psi$'s as non-abelian
generalizations of the classical theta functions.
Indeed, the latter are holomorphic sections of a line bundle
over the moduli space (the Jacobian) of the holomorhic
$\NC^*$-bundles over a Riemann surface.
\vskip 0.4cm
\subsection{States of the Chern-Simons theory}
Physically, the holomorphic maps $\Psi$ satisfying the Ward
identity (\ref{CSst}) may be identified as the quantum
states of the three-dimensional Chern-Simons (CS) gauge theory
\cite{27}. The classical phase space of the CS
theory on the 3-manifold $\,\Sigma\times\NR\,$ is composed
of the flat ${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$-valued gauge fields $\,i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A\,$ on $\Sigma$
modulo $G$-valued gauge transformations. The flatness
condition is
\begin{eqnarray}} \newcommand{\de}{\bar\partial
F(A)\,\equiv\,dA\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}+\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A^2\ =\ 0\,.
\label{fla}
\qqq
In the holomorphic quantization {\it \`{a} la} Bargmann,
the quantum states of the theory are described
as holomorphic functionals $\Psi$ on the space $\CA^{01}$
with the condition (\ref{fla}) imposed as a quantum constraint:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
F(A)\,\,\Psi(A^{01})\ =\ 0\,,
\label{flaq}
\qqq
with $F(A)$ as in Eq.\,\,(\ref{fla}) but with $A^{01}$ acting
as the multiplication operator and $A^{10}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$
as the differentiation:
$\,A^a_{z}=-\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{{2\pi}\over k}\,{{\delta}
\over{\delta A^a_{\bar z}}}$. \hspace{0.025cm}} \newcommand{\ch}{{\rm ch} The constraint (\ref{flaq})
is closely related to the infinitesimal Ward
identity:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
(\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} F(A)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}+\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{_{4\pi i}\over
^{k}}\sum\limits_n \delta_{\xi_{_n}}\,t^a\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} t^a_{_{R_n}})
\,\,\Psi(A)
\ =\ 0
\label{IWI}
\qqq
obtained by expanding the global Ward identity
(\ref{CSst}) to the first order in $h$ around $1$.
The infinitesimal identity (\ref{IWI}) is equivalent
to its global version (\ref{CSst}). In the absence
of insertions, it coincides with Eq.\,\,(\ref{flaq}).
The modifications involving the insertions correspond
in the CS gauge theory language to the insertions
of the Wilson lines $\{\xi_n\}\times\NR$ in
representations $R_n$.
\vskip 0.3cm
It is a crucial fact that the ($k$-dependent) spaces
$\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\NW_{_\Sigma}(\un{\xi},\un{R})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ of the holomorphic maps
$\Psi$'s satisfying the Ward identities (\ref{CSst})
or (\ref{IWI}) are finite-dimensional, with the dimension
given by the celebrated Verlinde formula \cite{Verl}.
In particular, only representations $R_n$ with
HW's $\lambda_{_{R_n}}$ integrable at level $k$ may
give rise to non-trivial spaces
$\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\NW_{_\Sigma}(\un{\xi},\un{R})$.
It is instructive to look more carefully into the case of
the Riemann sphere $\NC P^1$. \hspace{0.025cm}} \newcommand{\ch}{{\rm ch} On $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\NC P^1$ all gauge
fields $A^{01}$ may be written in the form $h^{-1}\de h$
or may be approximated by the fields of this form. In other
words, the gauge orbit of $A^{01}=0$ is dense in $\CA^{01}$.
But by Eq.\,\,(\ref{CSst}),
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\Psi(h^{-1}\de h)\ =\ \ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} S(h)}\
\mathop{\otimes}\limits_n h^{-1}_{_{R_n}}(\xi_n)\ \Psi(0)\,,
\label{etse}
\qqq
where $\,\Psi(0)\in\mathop{\otimes}\limits_nV_{_{R_n}}\,$
is an element of a finite-dimensional space. Hence $\Psi(0)$
determines $\Psi$ on a dense set of gauge fields $A$,
so everywhere. In fact, $\Psi(0)$ belongs to the subspace
$\,(\mathop{\otimes}\limits_nV_{_{R_n}})^{^G}\,$ of
tensors invariant under the diagonal action of $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} G$,
as is easy to see by taking constant $h$
in Eq.\,\,(\ref{etse}). We obtain then a natural embedding
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\NW_{_{\NC P^1}}(\un{\xi},\un{R})\ \ \subset\ \
(\mathop{\otimes}\limits_nV_{_{R_n}})^{^G}\,.
\qqq
The images of $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\NW_{_{\NC P^1}}(\un{\xi},\un{R})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$
in the spaces of invariant tensors are, in general, proper
subspaces of $\,(\mathop{\otimes}\limits_n
V_{_{R_n}})^{^G}$ depending on $k$. The reason
is that the $\Psi$'s defined
by Eq.\,\,(\ref{etse}) on $A^{01}=h^{-1}\de h$ do not
extend holomorphically to all of $\CA^{01}$
for all invariant tensors $\Psi(0)$.
In particular, the image of $\NW_{_{\NC P^1}}(\un{\xi},\un{R})$,
which is zero if some of HW's $\lambda_{_{R_n}}$ are not
integrable at level $k$, becomes the whole
space of invariant tensors for sufficiently large $k$.
\vskip 0.3cm
For genus one, i.e. on the complex tori
$\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{T}_\tau=\NC/(\NZ+\tau\NZ)$, a dense set of
gauge fields is formed by the gauge orbits
of the fields
\begin{eqnarray}} \newcommand{\de}{\bar\partial
A^{01}_{u}={_\pi\over^{{\rm Im}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\tau}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{u}\, d\bar z
\nonumber} \newcommand{\no}{\noindent
\qqq
with $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{u}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ in the complexified Cartan algebra ${{\bf t}}} \newcommand{\Nz}{{{\bf z}}^\NC
\subset{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}^\NC$. It is then enough to know the CS
states $\Psi$ only on the gauge fields $A^{01}_u$.
In particular, in the case with no insertions,
the holomorphic functions $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\psi\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ defined by
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\psi(u)\ =\ \ee^{-\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\pi\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} k\over2\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\rm Im}
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\tau}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\,\tr\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{u}^2}\,\,\Psi(A^{01}_{u})
\nonumber} \newcommand{\no}{\noindent
\qqq
characterize completely the CS states $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Psi\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$.
\vskip 0.3cm
It appears that the functions $\psi(u)$ are
arbitrary combinations of the characters
$\chi_{_{\hat R}}(\tau,\ee^{\m2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} u})$
of the HW representations of the current algebra $\hat{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$,
see Eq.\,\,(\ref{hatch}). This fact implies an important property
of the latter. Recall that the tori ${T}_\tau$
and ${T}_{\tau'}$ for $\tau'=-{1\over\tau}$ may be identified
by the map $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} z\mapsto z'=-z/\tau$. Under this
identification, $A^{01}_{u'}\mapsto A^{01}_u$ if $u'=-u/\tau$.
It follows then that the characters $\chi_{_{\hat{R}'}}(\tau',
\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} 2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} u'})$ of the current algebra
are combinations of the characters $\chi_{_{\hat{R}}}(\tau,
\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} 2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} u})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\chi_{_{\hat{R}'}}(\tau',\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} 2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} u'})\ =\
\sum\limits_{\hat{R}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} S_{_{R'}}^{^{R}}\,\,
\chi_{_{\hat{R}}}(\tau,\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} 2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} u})\,.
\nonumber} \newcommand{\no}{\noindent
\qqq
Hence the modular transformation $\tau\mapsto-1/\tau$
(and more generally, the transformations of $SL(2,\NZ)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$)
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} may be implemented on the characters of the
current algebra. The symmetric unitary matrices
$(S_{_{R'}}^{^{R}})$ representing the action
of the transformation $\tau\mapsto-1/\tau$ may be expressed
explicitly by the characters $\chi_{_R}$ of the
group $G$ and by the Weyl denominator of Eq.\,\,(\ref{WD}),
\begin{eqnarray}} \newcommand{\de}{\bar\partial
S^{^{R}}_{_{R'}}\ =\ 1^{_1\over^4}\,\vert T\vert^{-{_1\over^2}}\,\,
\chi_{_{R'}}(\ee^{\m2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\hat\lambda_{_{R}}/\hat k})\,\,
\Pi(\ee^{\m2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\hat\lambda_{_{R}}/\hat k})\ =\
S^{^{R'}}_{_{R}}\ =\ \overline{S^{^{\overline{R}}}_{_{R'}}}
\label{modm}
\qqq
in the notation: $\hat{\lambda}\equiv\lambda+\rho$ and
$\hat{k}\equiv k+h^\vee$ with $\rho$ the Weyl vector
and $h^\vee$ the dual Coxeter number.
The normalizing factor $\vert\hat T\vert$ is the number
of the Cartan group elements of the form
$\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} 2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\hat{\lambda}/\hat{k}}$ with $\lambda$
a weight. $1^{_1\over^4}$ is a fourth root of unity.
For the $SU(2)$ group, the above formula reduces to
\begin{eqnarray}} \newcommand{\de}{\bar\partial
S^{^{j}}_{_{j'}}\ =\ \left({_2\over^{k+2}}\right)^{_1\over^2}\,
\sin{_{\pi\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}(2j+1)(2j'+1)}\over^{k+2}}\,.
\qqq
\vskip 0.4cm
\subsection{Verlinde dimensions and the fusion ring}
The dimensions $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\hat N_{_{\un{R}}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ of the spaces
$\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\NW_{_\Sigma}(\un{\xi},\un{R})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ are independent of the
complex structure of $\Sigma$ and the locations $\un{\xi}$
of the insertion points (but dependent on the level
$k$ of the theory suppressed in the notation).
They are given by the Verlinde formula which, in the present
context, is a natural generalization of the classical
formula for the dimensions $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} N_{_{\un{R}}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ of the spaces
$\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}(\otimes V_{_{R_n}})^{^G}$ of group $G$ invariant tensors.
The dimensions $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} N_{_R}$ may be computed from the characters
of the representations $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} R_n$:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
N_{_{\un{R}}}\ =\ \int_{_G}\hspace{-0.02cm}\prod\limits_n
\chi_{_{R_n}}(g)\,\, dg
\ =\ \int\prod\limits_n\chi_{_{R_n}}(\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} 2\pi
i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda/k})
\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\,\,\vert\Pi(\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} 2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda/k})
\vert^2\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} d\lambda\,,
\label{wWD}
\qqq
where we have used the relation (\ref{WDcl}).
For simple, simply connected groups, the last integral
may be taken over the symplex
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\Delta_k\ \,=\,\ \{\,\lambda\in\Ntt\,\,\vert\,
\,\tr\,(\alpha^\vee\lambda)\geq 0
\ {\rm for}\ \alpha>0,\ \tr\,(\phi^\vee\lambda)\leq k\,\}
\label{sympl}
\qqq
whose elements label the conjugacy classes classes $\CC_\lambda$
in a one to one way.
Note that the weights in the symplex $\Delta_k$ are exactly
the HW's integrable at level $k$, see the definition (\ref{iHW}).
The numbers $N_{_{R\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} R_1\overline{R}_2}}$ coincide with
the dimensions $N_{_{R\, R_1}}^{^{R_2}}$ of the multiplicity
spaces in the decomposition (\ref{tpd}) of the tensor
product of representations, i.e. with the structure
constants of the character ring $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\cal R}} \newcommand{\CS}{{\cal S}_{_G}$ of the group $G$, see
Eq.\,\,(\ref{strf}).
For example for the $SU(2)$ group, $\,N_{_{j\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} j_1}}^{^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} j_2}}
=1\,$ if $\,\vert j-j_1\vert\leq j_2\leq j+j_1\,$ and $\,j+j_1+j_2\,$
is an integer and $\,N_{_{j\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} j_1}}^{^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} j_2}}=0\,$ otherwise.
The ring $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\cal R}} \newcommand{\CS}{{\cal S}_{_G}$ comes with
an additive $\NZ$-valued form $\omega$
given by the integral over $G$. $\,\omega$ assigns to the combination
$\sum n_i\chi_{_{R_i}}$ of characters the coefficient of the
character $\chi_{_1}=1$ of the trivial representation $R=1$.
The dimensions $N_{_{\un{R}}}$ are the values of $\omega$
on the product of the characters $\chi_{_{R_n}}$ in $\,{\cal R}} \newcommand{\CS}{{\cal S}_{_G}$.
\vskip 0.3cm
The dimensions $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\hat N_{_{\Sigma,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\un{R}}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ of the spaces
$\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\NW_{_\Sigma}(\un{\xi},\un{R})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ are given by the formula
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\hat N_{_{\Sigma,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\un{R}}}\ =\ {_1\over^{\vert\hat T\vert}}
\sum\limits_{{\rm weights}\atop\lambda\in\Delta_k}
\prod\limits_n\chi_{_{R_n}}(\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} 2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\hat{\lambda}/
\hat{k}})
\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\,\,\vert\Pi(\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} 2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\hat{\lambda}/\hat{k}})
\vert^{^{2-g_{_\Sigma}}}\,,
\label{wWDd}
\qqq
in the notations from the end of the last subsection
and with $g_{_\Sigma}$ denoting the genus of the surface
$\Sigma$. The above equation is
a rewrite of the original Verlinde formula:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\hat N_{_{\Sigma,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\un{\xi}}}\ =\ \sum\limits_{\hat R}
\prod\limits_n (S_{_{R_n}}^{^{R}}/S_{_1}^{^{R}})\
(S_{_1}^{^{R}})^{^{2-g_{_\Sigma}}}
\label{orV}
\qqq
which may be easily obtained from Eq.\,\,(\ref{wWDd})
with the use of the explicit expression (\ref{modm})
for the modular matrix $(S_{{R'}}^{R})$.
For the particular case of $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Sigma=\NC P^1$, Eq.\,\,(\ref{wWDd})
is clearly a deformation of Eq.\,\,(\ref{wWD}). More exactly,
the sum in Eq.\,\,(\ref{wWDd}) is a Riemann sum approximation
of the integral in Eq.\,\,(\ref{wWD}).
The genus zero 3-point dimensions $\,\hat N_{_{R\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} R_1\overline{R}_2}}
\equiv\hat N_{_{R\, R_1}}^{^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} R_2}}$ give the structure
constants of a commutative ring $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\hat{\cal R}} \newcommand{\CS}{{\cal S}}_{_G}$
which is additively generated by the representations $R$
with the HW's integrable at level $k$.
The ($k$-dependent) ring $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\hat{\cal R}} \newcommand{\CS}{{\cal S}}_{_G}$ is called
the {\bf fusion ring} of the WZW model.
For the $SU(2)$ group and all spins $\leq{k\over 2}$,
$\,{\hat N}_{_{j\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} j_1}}^{^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} j_2}}
=1\,$ if $\,\vert j-j_1\vert\leq j_2\leq j+j_1\,$
and $\,j+j_1+j_2\,$ is an integer $\leq k$ and
$\,{\hat N}_{_{j,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} j_1}}^{^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} j_2}}=0\,$ otherwise.
The fusion ring is a deformation
of the character ring $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\cal R}} \newcommand{\CS}{{\cal S}_{_G}$. More exactly,
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{\hat{\cal R}} \newcommand{\CS}{{\cal S}}_{_G}\ \cong\ {\cal R}} \newcommand{\CS}{{\cal S}_{_G}/\hat\CI\,,
\qqq
where $\hat\CI$ is the ($k$-dependent) ideal in ${\cal R}} \newcommand{\CS}{{\cal S}_{_G}$
composed of the functions vanishing on the Cartan group elements
$\,\ee^{\m2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\hat{\lambda}/\hat{k}}\,$ for weights
$\lambda\in\Delta_k$. The isomorphism identifies
the image of $\chi_{_R}$ in ${\cal R}} \newcommand{\CS}{{\cal S}_{_G}/\hat\CI$ with the generator
of $\,{\hat{\cal R}} \newcommand{\CS}{{\cal S}}_{_G}$ corresponding to $R$ for representations
$R$ with integrable HW's. The coefficient at the generator corresponding
to the trivial representation defines an additive $\NZ$-valued form
$\hat\omega$ on ${\hat{\cal R}} \newcommand{\CS}{{\cal S}}_{_G}$. For all $R_n$ with integrable HW's,
the genus zero Verlinde dimensions $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\hat N_{_{\un{R}}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$
are given by the values of $\hat\omega$
on the image in the fusion ring of the product
of the characters $\chi_{_{R_n}}$. For fixed
representations, $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\hat N_{_{R\, R_1}}^{^{R_2}}
=N_{_{R\, R_1}}^{^{R_2}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ for sufficiently high $k$.
The fusion ring may be also identified as the character
ring of the quantum deformation $\CU_q({{\bf g}}} \newcommand{\Ntt}{{{\bf t}})$ of the
enveloping algebra of ${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$ for $q=\ee^{\pi i/(k+h^\vee)}$,
an example of the intricate relations between the
WZW model and the quantum groups.
\vskip 0.4cm
\subsection{Holomorphic factorization}
Consider now the Ward identity (\ref{WIcfr}) for the mirror
chiral gauge transformations.
The anti-holomorphic maps $\Phi$ of the space $\CA^{10}$
of the ${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}^\NC$-valued 1,0-gauge fields $A^{10}$ with values
in $\mathop{\otimes}\limits_n{V_{_{\overline{R}_n}}}$ such that
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\Phi(A^{10})\ =\ \ee^{-S(h,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A^{10})}\
\mathop{\otimes}\limits_n{h^{\,t}_{_{\overline{R}_n}}(\xi_n)}\
\Phi({}^{h}\hspace{-0.09cm}A^{10})
\nonumber} \newcommand{\no}{\noindent
\qqq
are the complex conjugates of the holomorphic maps $\Psi$
satisfying the relation (\ref{CSst}):
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\Phi(A^{10})\ =\ \overline{\Psi(-(A^{10})^*)}\,,
\qqq
where the star denotes the anti-linear involution of the
complexified Lie algebra ${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}^\NC$ leaving ${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$ invariant
(it coincides with the hermitian conjugation for ${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}=su(N)$).
It follows that the correlation functions, in their
dependence of the external gauge field, are sesqui-linear
combinations of the elements of the space $\NW_{_\Sigma}
(\un{\xi},\un{R})$ of holomorphic solutions of
Eq.\,\,(\ref{CSst}):
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{\Sigma}}(A)\,<\mathop{\otimes}\limits_n g_{_{R_n}}
(\xi_n)>_{_{\hspace{-0.08cm}\Sigma}}\hspace{-0.1cm}(A)\
=\ H^{\alpha\beta}\,
\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Psi_\alpha(A^{01})\otimes\overline{\Psi_\beta(-(A^{01})^*)}\,,
\label{sesq}
\qqq
where the states $\Psi_\alpha$ form a basis
of $\NW_{_\Sigma}(\un{\xi},
\un{R})$ and the right hand side should be summed over
$\alpha$ and $\beta$. The equality involves the natural identification
of the vector spaces $\,(\mathop{\otimes}\limits_nV_{_{{R}_n}})
\otimes(\mathop{\otimes}\limits_nV_{_{{\overline{R}}_n}})
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\cong\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\mathop{\otimes}\limits_nEnd(V_{_{{R}_n}})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$.
\,The partition function ${\cal Z}} \newcommand{\s}{\hspace{0.05cm}(A)$ has to be given
by similar expressions pertaining to the case without
insertions. In particular, on the complex torus ${T}_\tau$,
and in the vanishing gauge field,
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{{T}_{_\tau}}}\ =\ \sum\limits_{\hat R,\,\hat {R}'}
H^{^{\hat R\,\hat{R}'}}\ \ch_{_{\hat R}}(\tau,1)\,\,
\overline{\ch_{_{\hat{R}'}}(\tau,1)}\,.
\label{tpf0}
\qqq
\vskip 0.3cm
The matrices $(H^{\alpha\beta})$ should be specified for
a given choice of the basis $(\Psi_\alpha)$ for each complex
structure on the surface and for each configuration of the insertions
points so that, if we did not have means to compute them,
the above formulae would mean little progress
towards the solution of the WZW theory. Fortunately,
there exist effective ways to determine the coefficients
$H^{\alpha\beta}$.
\vskip 0.5cm
\subsection{Scalar product of the Chern-Simons states}
It was argued in \cite{41}, see also
\cite{Witten} for a formal functional integral argument,
that the matrices $(H^{\alpha\beta})$ appearing
in Eq.\,\,(\ref{sesq}) are inverse to the matrices
$(H_{\alpha\beta})$ with matrix elements
\begin{eqnarray}} \newcommand{\de}{\bar\partial
H_{\alpha\beta}\ =\ (\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Psi_\alpha,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Psi_\beta\hspace{0.025cm}} \newcommand{\ch}{{\rm ch})
\nonumber} \newcommand{\no}{\noindent
\qqq
given by the scalar product of the CS states.
According to the rules of holomorphic
quantization, the latter is given by the functional integral
\begin{eqnarray}} \newcommand{\de}{\bar\partial
(\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Psi,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Psi'\hspace{0.025cm}} \newcommand{\ch}{{\rm ch})\ =\ \int\limits_{\CA^{01}}
(\Psi(A^{01}),\Psi'(A^{01})
)_{_{\otimes V_{_{R_n}}}}\,
\ee^{-\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{k\over 2\pi}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Vert A\Vert^2}
\ DA
\label{gfin}
\qqq
over the ${{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$-valued gauge fields $\,i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A=i(-(A^{01})^*+A^{01})\,$
with \,$\Vert A\Vert^2\equiv i\int_{_\Sigma}\hspace{-0.05cm}
\tr\,(A^{01})^*\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A^{01}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}.$ \,This is again a formal
expression. The point is, however, that the $DA$-integral
may be calculated exactly by reducing it to doable
Gaussian (i.e. free field) functional integrals.
Ones this is done, the exact solution for the correlation
functions follows by Eq.\,\,(\ref{sesq}). Note that
the above solution for coefficients $H^{\alpha\beta}$
guarantees that the right hand side of Eq.\,\,(\ref{sesq})
is independent of the choice of a basis of the CS
states. Let us briefly sketch how one achieves the reduction
of the integral (\ref{gfin}) to the free field ones.
\vskip 0.3cm
In the first step, the integral (\ref{gfin}) may be rewritten
by a trick resembling the Faddeev-Popov treatment of
gauge theory functional integrals. The reparametrization of
the gauge fields
\begin{eqnarray}} \newcommand{\de}{\bar\partial
A^{01}\s=\s{}^{h^{-1}}\hs{-0.11cm}A^{01}(n)
\label{chva}
\qqq
by the chiral $G^\NC$-valued gauge transforms of a (local) slice
$\,n\mapsto A^{01}(n)\,$ in the space $\,\CA^{01}\,$ cutting each
gauge orbit in one point, see Fig.\,\,10,
\leavevmode\epsffile[-90 -20 327 160]{fig10.eps}
\noindent permits to rewrite the functional
integral expression for the norm squared of a CS
state in the new variables as
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\Vert\Psi\Vert^2\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}=\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\int(\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Psi(A^{01}(n),
\otimes(hh^*)_{_{R_n}}^{-1}
\Psi(A^{01}(n)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch})_{_{V_{_{_{R_n}}}}}\hspace{-0.07cm}
\ee^{(k+2h^\vee)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} S(hh^*,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A^{01}(n))}\ D(hh^*)
\ d\mu(n)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}.\hs{0.7cm}
\label{HWZW}
\qqq
To obtain the expression on the right hand side,
we have used Eq.\,\,(\ref{CSst}). The term $\,2\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} h^\vee\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} S(hh^*)\,$
in the action comes from the Jacobian of the change of variables
(\ref{chva}) contributing also to the measure $\,d\mu(n)\,$
on the local slice in $\CA^{01}$.
\vskip 0.3cm
Unlike in the standard Faddeev-Popov setup, the integral
over the group of gauge transformations did not
drop out since the integrand in Eq.\,\,(\ref{gfin}) is invariant
only under the $\,G$-valued gauge transformations. Instead,
we are left with a functional integral (\ref{HWZW})
similar to the one for the original
correlation functions, see Eq.\,\,(\ref{cfu}), except
that it is over fields $\,hh^*\,$. These fields may be considered
as taking values in the contractible hyperbolic space
$\,G^\NC/G\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$. $\,D(hh^*)\,$ is the formal product
of the $\,G^\NC$-invariant measures on $\,G^\NC/G\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$. \,The
gain is that the functional integral (\ref{HWZW}) for the
hyperbolic WZW model correlation functions is doable.
For example for $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} G=SU(2)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ and for $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Sigma=\NC P^1$,
where we may set $A^{01}(n)=0$ (in this case the gauge orbit
of $A^{01}=0$ gives already a dense open subset $\CA^{01}$),
\begin{eqnarray}} \newcommand{\de}{\bar\partial
S(hh^*)\ =\ -\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{_i\over^{2\pi}}\int\partial} \newcommand{\ee}{{\rm e}\phi\wedge\de\phi\s
-\s{_i\over^{2\pi}}\int(\partial} \newcommand{\ee}{{\rm e}+\partial} \newcommand{\ee}{{\rm e}\phi)\bar v\wedge(\de+\de\phi)v
\nonumber} \newcommand{\no}{\noindent
\qqq
in the Iwasawa parametrization
\s\s$
h\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}=\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}({_{\ee^{\phi/2}}\atop^0}\,{_0\atop^{\ee^{-\phi/2}}})
\,({_1\atop^0}\,{_v\atop^1})\, u\,\,$ of the 3-dimensional
hyperboloid $\,SL(2,\NC)/SU(2)\,$ by $\,\phi\in\NR\,$ and
$\,v\in\NC\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$, \hspace{0.025cm}} \newcommand{\ch}{{\rm ch} with $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} u\in SU(2)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$. \,Although the action
is not quadratic in the fields, it is quadratic
in the complex field $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} v$ so that the $v$ integral
can be explicitly computed. Somewhat miraculously,
the resulting integral appears to depend on the remaining
field $\phi$ again in a Gaussian way
so that the integration may be carried out further.
The same happens for more complicated groups and on surfaces
with handles, except that the procedure requires more steps.
At the end, one obtains explicit finite-dimensional
integrals. Hence, the integral (\ref{HWZW}) belongs
to the class of functional integrals that may be explicitly
evaluated. The Gaussian functional integrals encountered
in its computation require mild renormalizations
(the zeta-function or similar regularization of determinants,
Wick ordering of insertions) but these are well
understood. They are responsible for the mild non-invariance
of the WZW correlation functions under the local rescalings
of the metric, leading to the values of the Virasoro
central charge and of the conformal dimensions discussed
above.
\vskip 0.3cm
On the complex tori ${T}_\tau$ with no insertions and in
the constant metric $|dz|^2$, the scalar product of
the CS states takes a particularly simple form:
the current algebra characters $\chi_{_{\hat R}}(\tau,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\cdot\hspace{0.025cm}} \newcommand{\ch}{{\rm ch})$
appear to give an orthonormal basis of the space
of states. It follows that the toroidal partition function
in the constant metric is
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{\cal Z}} \newcommand{\s}{\hspace{0.05cm}(\tau)\ =\ \sum\limits_{\hat R}|\chi_{_{\hat R}}(\tau,1)|^2\,,
\label{diag}
\qqq
see Eq.\,\,(\ref{tpf0}). The exact normalization
of the constant metric on ${T}_\tau$ is not important
since at genus one constant rescalings of the metric, exceptionally,
do not effect the partition functions. The latter fact has
an important consequence. It implies that the partition function
${\cal Z}} \newcommand{\s}{\hspace{0.05cm}(\tau)$ has to be a modular invariant:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{\cal Z}} \newcommand{\s}{\hspace{0.05cm}(\tau)\ =\ {\cal Z}} \newcommand{\s}{\hspace{0.05cm}({_{a\tau+b}\over^{c\tau+d}})
\nonumber} \newcommand{\no}{\noindent
\qqq
for $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}({_a\atop^c}\,{_b\atop^d})\in SL(2,\NZ)$.
This is indeed the property of the right hand side
of Eq.\,\,(\ref{diag}) since the matrices implementing
the modular transformations on the characters of the current algebra
are unitary.
\vskip 0.3cm
Explicit finite-dimensional integral formulae for the scalar
product (\ref{gfin}) have been obtained for general
groups at genus zero and one and, for \,$G=SU(2)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$, \hspace{0.025cm}} \newcommand{\ch}{{\rm ch} for higher
genera. It is clear that the case of general group and genus $>$1
could be treated along the same lines, but the explicit calculation
has not been done. It should be also said that the general
proof of the convergence of the resulting finite-dimensional
integrals is also missing, although several special cases
have been settled completely.
\vskip 0.4cm
\subsection{Case of $\,G=SU(2)$ at genus zero}
To give a feeling about the form of the explicit
expressions for the scalar product of the CS
states, let as describe the result for $G=SU(2)$
and $\Sigma=\NC P^1$ with insertions at points
$z_n$ in the standard complex coordinate $z$.
In this case, as we have discussed above,
the CS states correspond to invariant
tensors $v$
in $\,(\mathop{\otimes}\limits_nV_{_{j_n}})^{^{SU(2)}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$,
\,where we label the irreducible representations of $SU(2)$
by spins. The spin $j$ representation acts
in the space $\,V_{_j}\,$ spanned by the vectors
$\,(\sigma_-)_{_j}^\ell\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
v_{_{j}}\,$ with $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\ell=0,1,\dots,2j\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},$ where
$\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} v_{_j}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ is the normalized HW vector annihilated
by $\,(\sigma_+)_{_j}$, with $\sigma_i$ denoting the Pauli
matrices. For the scalar product of the CS
states, the procedure described in the previous subsection
gives the following integral expression:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\Vert v\Vert^2\ =\ f(\sigma,{\un z},{\un j},k)
\int\limits_{\NC^J}\Big\vert\,(\, v\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},\s\omega(\un{z},\un{y})
\,)_{_{\otimes V_{_{R_n}}}}
\,\ee^{-{2\over k+2}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} U(\un{z},\un{y})}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Big\vert^2\,\,
\prod\limits_{a=1}^Jd^2y_a\,.
\label{scP0}
\qqq
Let us explain the terms on the right hand side. First,
\begin{eqnarray}} \newcommand{\de}{\bar\partial
f(\sigma,{\un{z}},{\un{j}},k)\ =\
\ee^{-{1\over 2\pi i(k+2)}\int\partial} \newcommand{\ee}{{\rm e}\sigma\de\sigma}
\,\left({_{{\det}'(-\Delta)}\over^{{area}}}
\right)^{\hspace{-0.06cm}3/2}
\prod\limits_n\ee^{\m2\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{j_n(j_n+1)\over k+2}
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\sigma(z_n)}
\label{explf}
\qqq
carries the dependence on the metric
\s$\ee^{2\sigma}\vert dz\vert^2\,$
on $\,\NC P^1\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$, \,with $\,{\rm det}'(-\Delta)\,$ denoting
the zeta-function regularized determinant of the
(minus) Laplacian on $\NC P^1$ with omission of
the zero eigenvalue. The $\sigma$-dependence of
$({\rm det}'(-\Delta)/{area})$ is given by a term
$\,\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{1\over4\pi i}\int\partial} \newcommand{\ee}{{\rm e}\sigma\de\sigma}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$
leading altogether to the value ${3\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} k\over k+2}$ of
the Virasoro central charge of the WZW theory (recall
that this is the inverse of the scalar product
that enters the WZW correlation functions).
Similarly, the conformal weight $\Delta_j$ of the
fields $g_j(\xi)$ of the $SU(2)$ WZW theory may be
read from Eq.\,\,(\ref{explf}) to be ${j(j+1)\over k+2}$.
\,Next, $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\omega(\un{z},\un{y})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ is a meromorphic
$\,\mathop{\otimes}\limits_nV_{_{j_n}}$-valued function
of $\un{z}$ and $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\un{y}=(y_1,\dots,y_{_J})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$, \hspace{0.025cm}} \newcommand{\ch}{{\rm ch} where
$J=\sum_{_n}\hspace{-0.05cm}j_n\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\omega(\un{z},\un{y})\ =\ \prod\limits_{r=1}^J\Big(\sum\limits_n
{_1\over^{y_r-z_n}}\s (\sigma_-)_{_{j_n}}
\Big)\,\mathop{\otimes}\limits_n v_{_{j_n}}\,.
\qqq
Finally, $\,U(\un{z},\un{y})\s$ is a multivalued function
\begin{eqnarray}} \newcommand{\de}{\bar\partial
U(\un{z},\un{y})\ =\ \sum\limits_{n<m}
j_n j_{m}\s\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\ln(z_n-z_m)
\s-\s\sum\limits_{n,r}j_n\s\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\ln(z_n-y_r)\s+\s\sum\limits_{r<s}
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\ln(y_r-y_s)\,.
\qqq
\vskip 0.3cm
The integral (\ref{scP0}) is over a positive density with
singularities at the coinciding $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} y_r\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$'s and the question
arises as to whether it converges.
A natural conjecture states that the integral
is convergent if and only if the invariant
tensor $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} v\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ is in the image of the space of
states $\,\NW(\NC P^1,\un{z},\un{j})$ explicitly
described as the set of $\,v\in(\mathop{\otimes}
\limits_nV_{_{j_n}})^{^{SU_2}}\,$ such that
\begin{eqnarray}} \newcommand{\de}{\bar\partial
(\otimes v_{j_n}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},\,\prod\limits_{n}
(\sigma_+)_{_{j_n}}^{p_n}
\,\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} z_n\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}(\sigma_+)_{_{j_n}}}\s v\hspace{0.025cm}} \newcommand{\ch}{{\rm ch})_{_{\otimes V_{_{j_n}}}}
\hspace{-0.15cm}=\,0\ \,\ \quad{\rm for}\quad\ \sum p_n
\leq J-k-1\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}.
\qqq
In particular, for two or three points, the image does not
depend on the location of the insertions and gives the whole
space of invariant tensors if $\sum j_n\leq k$ and zero
otherwise. In this case, the integrals in Eq.\,\,(\ref{scP0})
may indeed be computed in a closed form confirming
the conjecture. Numerous other special cases have been
checked. In general, the ``only if'' part of the conjecture
is easy but the ``if'' part remains to be verified.
\vskip 0.4cm
\subsection{Knizhnik-Zamolodchikov connection}
There is another way to construct the matrices
$(H^{\alpha\beta})$ entering the formula (\ref{sesq}) for
the correlation functions. Let us describe it briefly.
\vskip 0.3cm
The spaces $\NW_{_\Sigma}(\un{\xi},\un{R})$ of the CS
states depend on the complex structure of the surface $\Sigma$
and on the insertion points. They form, in a natural way,
a holomorphic vector bundle $\CW$ whose base is the space of
complex structures on a given smooth surface $\Sigma$ and of
non-coinciding insertions $\un{\xi}$ (modulo
diffeomorphisms). The scalar product of the CS
states equips this bundle with a hermitian
structure. In turn, a hermitian structure on a holomorphic
vector bundle determines a unique connection
such that the covariant derivatives of the structure
and of the holomorphic sections vanish. Although the scalar
product of the CS states has been rigorously
defined in the general case only modulo the convergence
of finite-dimensional integrals (see the end of the
last subsection), the connection on the bundle
of the states may be easily constructed with full
mathematical rigor. It appears to be projectively
flat (i.e. with a curvature that is a scalar 2-form
on the base space), a crucial fact. In other words,
the parallel transport of a CS state around a closed loop
in the space of complex structures and insertions at most
changes its normalization.
\vskip 0.3 cm
For the genus zero case, there is only one complex
structure modulo diffeomorphisms.
If we fix it as the standard complex structure
on $\NC P^1$, then we are only left with the freedom to move
the insertion points $\un{z}$. The bundle $\CW$ is in this
case a subbundle of the trivial bundle with the fiber
formed by the invariant group tensors
$(\mathop{\otimes}\limits_nV_{_{R_n}})^{^G}$
and the connection may be extended to this trivial bundle.
The covariant derivatives of the sections of the latter
are given explicitly by the equations:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\nabla_{_{\bar z_n}}v\ =\ \partial} \newcommand{\ee}{{\rm e}_{_{\bar z_n}}v\s,\quad\quad
\nabla_{_{z_n}}v\ =\
\partial} \newcommand{\ee}{{\rm e}_{_{z_n}}v\s+\s{_2\over^{k+h^\vee}}\sum\limits_{m\not= n}
{_{t^a_{_{R_m}}\hspace{-0.09cm}t^a_{_{R_n}}}\over^{z_m-z_n}}\,v
\label{KZ}
\qqq
which go back to the work \cite{5} of Knizhnik-Zamolodchikov
on the WZW theory. In fact the above connection is
flat as long as the insertion points stay away from
infinity and the article \cite{5} studied their
horizontal sections such that $\nabla v=0$.
The higher genus generalizations of these equations
were first considered by Bernard \cite{10}\cite{11}
and we shall call the connection on the bundle $\CW$
the Knizhnik-Zamolodchikov (KZ) connection
for genus zero or the Knizhnik-Zamolodchikov-Bernard (KZB)
one for higher genera.
\vskip 0.3cm
In general, the KZB connection can be made flat by some
choices (as in the case of genus zero, where the curvature
has been concentrated at infinity). For a flat connection,
one may choose locally a basis $(\Psi_\alpha)$ of
horizontal sections. The gain from such
a choice of the basis of the CS
states is that the coefficients $H^{\alpha\beta}$
in Eq.\,\,(\ref{sesq}) become then independent
of the complex structure or the positions of the insertions.
Indeed, since $H_{\alpha\beta}
=(\Psi_\alpha,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Psi_\beta)$ and the KZB connection
preserves the scalar product of the states, the
above scalar products are constant for horizontal
$\Psi_\alpha$. Since the horizontal sections are, in particular,
holomorphic, Eq.\,\,(\ref{sesq}) gives then
a holomorphic factorization of the correlation functions
into sesqui-linear combinations of the {\bf conformal blocks}
holomorphic in their dependence on the complex structure
and positions of insertions. Such a finite factorization
is the characteristic feature of {\bf rational} CFT's.
As we see, the conformal blocks of the WZW theory
are given by the horizontal sections $\Psi_\alpha$
of the bundle $\CW$ of the CS states.
For example, at genus zero with no insertions,
the conformal blocks are formed by the characters
of the current algebra and Eq.\,\,(\ref{diag})
provides a particular realization of the holomorphic
factorization.
\vskip 0.3cm
Since the KZB connection, although flat, has nevertheless
a non-trivial holonomy, the conformal blocks are, in general,
multivalued. The coefficients $H^{\alpha\beta}$
in Eq.\,\,(\ref{sesq}) may then be fixed, up to normalization,
by demanding that the correlation
functions be uni-valued. The overall
normalization may be fixed, in turn, by considering
the limits when the insertion points coincide.
This was the strategy used in the original work \cite{5}
to compute the 4-point correlation function of the spin
${_1\over^2}$ field $g_{{_1\over^2}}(\xi)$ of the $SU(2)$ WZW model
on the Riemann sphere. The horizontality relations
for the conformal blocks reduce in this case to
the hypergeometric equation and the calculations
of the conformal blocks and of their monodromy are easy
to perform. For general genus-zero conformal blocks, one
obtains generalizations of the hypergeometric equation
whose solutions may be expressed by contour integrals
\cite{14}\cite{40}. The latter are, essentially,
the holomorphic versions of the integrals (\ref{scP0}) so
that the two strategies to obtain explicit solutions for
the correlation functions, one based on the study
of the monodromy of the conformal blocks and the other
one involving a calculation of the scalar product of the
CS states, are closely related.
\vskip 0.3cm
There appears to be a very rich structure behind
the connection (\ref{KZ}) and its generalizations.
It is closely related to the integrable systems of
mechanics and statistical mechanics, see
e.g.\,\,\cite{GaTr}\cite{FrResh}.
The holonomy of the connection gives representations
of the braid groups which played an important role in
the construction of the Jones-Witten invariants of knots
and 3-manifold invariants \cite{27}. The perturbative
solutions of the horizontality equations enter the Vasiliev
invariants of knots \cite{Konts}. The KZ connection is also
closely connected to quantum groups \cite{Feldqg}
and to Drinfel'd's quasi-Hopf algebras \cite{Drinf},
to mention only some interrelated topics.
\vskip 0.5cm
\nsection{Coset theories}
There is a rich family of CFT's
which may be obtained from the WZW models by a simple
procedure known under the name of a {\bf coset construction}
\cite{42}\cite{43}. On the functional integral level,
the procedure consists of coupling the \s$G$-group WZW
theory to a gauge field $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} B=i(B^{10}+\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} B^{01})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$
with values in a subalgebra
$\,{{\bf h}}} \newcommand{\Nr}{{{\bf r}}\subset{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}\,$. The field $B$ is then integrated over with
gauge-invariant insertions \cite{45}\cite{46}\cite{47}\cite{49}.
Let $H\subset G$ be the subgroup of $G$ corresponding to ${{\bf h}}} \newcommand{\Nr}{{{\bf r}}$.
Choose elements $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} t_n\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ in the space $\,(End(V_{_{R_n}},
V_{_{r_n}}))^H\,$ of the intertwiners of the action of $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} H\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$
in the irreducible $\,G$- and $\,H$-representation spaces.
The simplest correlation functions of the $\,G/H\,$ coset
theory take the form
\begin{eqnarray}} \newcommand{\de}{\bar\partial
&&<\prod\limits_{i=1}^n \tr_{_{V_{_{r_n}}}}\hs{-0.1cm}(t_n\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
g_{_{R_n}}\hspace{-0.06cm}(x_n)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} t_n^\dagger)
>_{_{\hspace{-0.08cm}\Sigma}}\cr
&&\hspace{2cm}=\ {_1\over^{{{\cal Z}} \newcommand{\s}{\hspace{0.05cm}^{^{G/H}}_{_\Sigma}}}}
\int\,\prod\limits_{n}
\tr_{_{V_{_{r_n}}}}\hs{-0.1cm}(t_n\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g_{_{R_n}}\hspace{-0.06cm}
(x_n)\, t_n^\dagger)\ \hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\ee^{-k\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} S(g,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} B)-
{k\over 2\pi}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Vert B\Vert^2}\,\, Dg\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} DB
\,,\hspace{0.8cm}
\label{cfGH}
\qqq
where $\,{\cal Z}} \newcommand{\s}{\hspace{0.05cm}^{^{G/H}}_{_\Sigma}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
=\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\int\ee^{-k\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} S(g,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} B)-{k\over 2\pi}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Vert B\Vert^2}
\,Dg\, DB\,$ is the partition function of the $G/H$ theory.
Note that the $\,g$-field integrals are the ones of the
WZW theory and are given by Eq.\s\s(\ref{sesq}).
Consequently,
\begin{eqnarray}} \newcommand{\de}{\bar\partial
&&{\cal Z}} \newcommand{\s}{\hspace{0.05cm}^{^{G/H}}_{_\Sigma}\,<\prod\limits_{n}
\tr_{_{V_{_{r_n}}}}\hs{-0.1cm}(t_n\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g_{_{R_n}}\hspace{-0.06cm}
(x_n)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} t_n^\dagger)
>_{_{\hspace{-0.08cm}\Sigma}}\cr
&&\hspace{2cm}=\ H^{\alpha\beta}\int(\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\otimes t_n
\Psi_\beta(B^{01})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},\,\otimes t_n\Psi_\alpha
(B^{01})\,)_{_{\otimes V_{_{r_n}}}}\,
\ee^{-{k\over 2\pi}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Vert B\Vert^2}
\,\,DB\,.\hs{0.4cm}
\label{aneq}
\qqq
The composition with $\,\otimes t_n\,$ defines a linear map $T$
between the spaces of the group $G$ and the group $H$
CS states, i.e. $T:\NW_{_\Sigma}(\un{\xi},\un{R})\rightarrow
\NW_{_\Sigma}(\un{\xi},\un{r})\,$ with
\begin{eqnarray}} \newcommand{\de}{\bar\partial
(T\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} \Psi)(B^{01})\ =\ \mathop{\otimes}\limits_n t_n\,\Psi(B^{01})\,.
\qqq
Indeed, it is straightforward to check that the right hand
side satisfies the the group $H$ version of the Ward identity
(\ref{CSst}). Eq.\s\s(\ref{aneq}) may be rewritten with the
use of the map $T$ as
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{\cal Z}} \newcommand{\s}{\hspace{0.05cm}^{^{G/H}}_{_\Sigma}<\prod\limits_{n}
\tr_{_{V_{_{r_n}}}}\hspace{-0.1cm}(t_n\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g_{_{R_n}}
\hspace{-0.06cm}(x_n)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} t_n^\dagger)
>_{_{\hspace{-0.08cm}\Sigma}}\
=\ H^{\alpha\beta}\ (\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} T\Psi_\beta\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},\, T\Psi_\alpha\hspace{0.025cm}} \newcommand{\ch}{{\rm ch})
\ =\ \tr_{_{\NW_{_\Sigma}(\un{\xi},\un{R})}}\hspace{-0.05cm}
T^\dagger T\,.
\label{GHcf}
\qqq
Let $\,(T^\mu_\alpha)\,$ denote the (``branching'') matrix
of the linear map $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} T\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ in the bases
$\,(\Psi_\alpha)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ and $\,(\psi_\mu)\,$ of, respectively,
$\,\NW_{_\Sigma}(\un{\xi},\un{R})\,$ and
$\,\NW_{_\Sigma}(\un{\xi},\un{r})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$, i.e.
$T\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Psi_\alpha=T^\mu_\alpha\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\psi_\mu\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$. \,Then
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{\cal Z}} \newcommand{\s}{\hspace{0.05cm}^{^{G/H}}_{_\Sigma}<\prod\limits_{n}
\tr_{_{V_{_{r_n}}}}\hspace{-0.1cm}(t_n\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g_{_{R_n}}
\hspace{-0.06cm}(x_n)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} t_n^\dagger)
>_{_{\hspace{-0.08cm}\Sigma}}\
=\ H^{\alpha\beta}\,\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\overline{T^\mu_\beta}\,\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
h_{\mu\nu}\,\, T^\nu_\alpha\,,
\nonumber} \newcommand{\no}{\noindent
\qqq
where $\,h_{\mu\nu}=(\psi_\mu,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\psi_\nu)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$.
\,Since the above relations hold also for the partition function,
it follows that the calculation of the coset theory
correlation functions (\ref{cfGH}) reduces to that of the scalar
products of group \s$G\s$ and group \s$H\s$ CS
states, both given by explicit, finite-dimensional integrals.
\vskip 0.3cm
Among the simplest examples of the coset theories
is the case with the group $\,G=SU(2)\times SU(2)\,$
with level $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}(k,1)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ (for product groups, the levels may
be taken independently for each group) and with $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} H\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$
being the diagonal $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} SU(2)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ subgroup. The resulting
theories coincide with the unitary {\bf minimal series} of
CFT's with the Virasoro central charges
$\,c=1-{6\over{(k+2)(k+3)}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$, \hspace{0.025cm}} \newcommand{\ch}{{\rm ch} first considered by
Belavin-Polyakov-Zamolodchikov \cite{3}.
The Hilbert spaces of these theories are built
from irreducible unitary representations
of the Virasoro algebras with $\,0<c<1\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$.
The simpliest one of them with $\,k=1\,$ and $\,c={1\over2}\,$
describes the continuum limit of the Ising model at critical
temperature or the scaling limit of the massless $\,\phi^4_2\,$
theory. In particular, in the continuum limit the spins in
the critical Ising model are represented
by fields \s$\tr\,g_{{_1\over^2}}(\xi)\s$ where \s$g\s$ takes
values in the first \s$SU(2)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$. \s The corresponding correlation
functions may be computed as above. One obtains this way
for the 4-point function on the complex plane (or the Riemann
sphere) an explicit expression in terms of the hypergeometric
function.
\vskip 0.3cm
The $\,G/H\,$ coset theory with $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} H=G\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ is a prototype
of a two-dimensional topological field theory. As follows from
Eq.\,\,(\ref{GHcf}), the correlation functions of the fields
$\,\tr\, g_{_R}(\xi)\,$ are equal to the dimension of the
spaces $\,\NW_{_\Sigma}(\un{\xi},\un{R})\,$, normalized
by the dimension of \s$\NW_{_\Sigma}(\emptyset,\emptyset)\,$
and are given by the Verlinde formula (\ref{wWDd}).
In particular, they do not depend on the position
of the insertion points, a characteristic feature
of the correlation functions in topological field theories.
\vskip 0.6cm
\nsection{Boundary conditions in the WZW theory}
Discussing in Sects.\,\,3.4 and 4.1 above the classical
and the quantum amplitudes for the WZW model on surfaces
with boundary, we have admitted an arbitrary behavior
of the fields on the boundary. In physical situations,
one often has to constrain this behavior by imposing
the boundary conditions (BC) on the fields.
The simplest example is provided
by the Dirichlet or Neumann BC's for the free
fields which fix to zero, respectively, the tangent or
the normal derivative of the field (the absorbing
versus the reflecting condition). Such conditions leave
unbroken an infinite-dimensional set of symmetries of
the free field theory. We shall be interested in BC's
in the WZW model with a similar property.
\vskip 0.3cm
The theory of boundary CFT's was pioneered
by Cardy \cite{Cardy} and Cardy-Lewellen \cite{CardLew}.
It found its applications e.g.\,\,in the theory of
isolated impurities in condensed matter physics
(in the so called Kondo problem, a traditional
playground for theoretical ideas) \cite{Affleck}.
In string theory, the use of the Neumann BC for free
open strings has a long tradition \cite{GWSch}.
The realization that one should also consider
free open strings with the Dirichlet BC came much later
and gave rise to a theory of Dirichlet-
or D-{\bf branes} \cite{Polch}: the end of an open string,
some of whose coordinates are restricted by the Dirichlet BC,
moves on a surface (brane) of a lower dimension.
The D-branes provide the basic tool in the analysis
of the non-perturbative effects in string theory:
of the stringy solitons and of the strong-weak coupling
dualities \cite{Polch}. The general theory
of boundary CFT's is slowly becoming an important
technique of string theory (see e.g.\,\,\cite{ReckSch}).
\,Here, for the sake of illustration,
we shall discuss a particular class of BC's for the WZW
theory. These conditions constrain the group $G$ valued
field $g$ to stay over the boundary components in fixed
conjugacy classes of $G$. Such BC's were discussed in
\cite{AlSch}, see also \cite{KlimSev}. They clearly generalize
the Dirichlet BC of the free fields, contrary to the claim
in \cite{AlSch} (based on the conventions of reference
\cite{Okada}) associating them to the Neumann BC. Our
presentation, along similar lines, clarifies, hopefully,
some of the discussions of the above papers.
\vskip 0.5cm
\subsection{The action}
As before, we shall represent a Riemann surface
$\Sigma$ with boundary as $\Sigma'\setminus(\mathop{\cup}
\limits_mD_{m}^{^{^{{\hspace{-0.21cm}}o}}})$,
where $D_{m}$ are disjoint unit discs embedded in a closed
surface $\Sigma'$ without boundary,
see Fig.\,\,6. We have seen in Sect.\,\,3.4 that the classical
amplitudes $\,\ee^{-S(g)}\,$ of the fields
$\,g:\Sigma\rightarrow G\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ of the WZW model take
values in a line bundle $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\cal L}} \newcommand{\CM}{{\cal M}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ rather than being numbers.
The line bundle ${\cal L}} \newcommand{\CM}{{\cal M}$ over the loop group $LG$ is not trivial but
it may be trivialized over certain subsets of $LG$, for
example the ones formed by the loops taking values
in special conjugacy classes $\CC_{\lambda}$.
It will then be possible to give numerical values to
the amplitudes $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\ee^{-S_{_{\Sigma}}(g)}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ for
$\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g:\Sigma\rightarrow G\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ satisfying the BC's
\begin{eqnarray}} \newcommand{\de}{\bar\partial
g(\partial} \newcommand{\ee}{{\rm e} D_{m})\ \subset\ \CC_{\lambda_m}\,.
\label{bc}
\qqq
In order to achieve this goal, we shall fix the 2-forms
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\omega_{_\lambda}\ =\
\tr\,(g^{-1}dg)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}(1-Ad_g)^{-1}(g^{-1}dg)\ =\
\tr\,\,(g_0^{-1}dg_0)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
\ee^{\m2\pi i\lambda/k}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}(g_0^{-1}dg_0)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
\ee^{-2\pi i\lambda/k}
\label{om}
\qqq
on the conjugacy classes $\,\CC_{\lambda}$ composed of the
elements $\,g\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}=\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g_0\,\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} 2\pi i\lambda/k}g_0^{-1}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$
(the operator $(1-Ad_g)$ is invertible on the vectors tangent
to $\CC_{\lambda}$). It is easy to check by a direct
calculation that $\,d\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\omega_{_\lambda}\,$ coincides
with the restriction of the 3-form $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\chi={1\over 3}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\tr
\,(g^{-1} dg)^3\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ to $\CC_\lambda$.
\vskip 0.3cm
Since the conjugacy classes in a simply connected group $G$
are simply connected, any field $g:\Sigma\rightarrow G$
satisfying the BC's (\ref{bc})
may be extended to a field $g':\Sigma'\rightarrow G$
in such a way that $g'(D_{m})\subset\CC_{\lambda_m}$
and then to a field $\widetilde} \def\hat{\widehat g$ on a 3-manifold $B$
such that $\partial} \newcommand{\ee}{{\rm e} B=\Sigma'$, see Fig.\,\,11.
\leavevmode\epsffile[-20 -20 367 157]{fig11.eps}
\noindent Having done this,
we define the Wess-Zumino part of the action as
\begin{eqnarray}} \newcommand{\de}{\bar\partial
S^{^{WZ}}_{\Sigma}(g)\ =\ {_k\over^{4\pi i}}
\int_{_{B}}\hspace{-0.08cm}\widetilde} \def\hat{\widehat g{}^*\chi\
-\ {_k\over^{4\pi i}}\sum\limits_m
\int_{_{D_{_m}}}\hspace{-0.1cm}\widetilde} \def\hat{\widehat g|_{_{D_{_m}}}^{\, *}
\omega_{_{\lambda_m}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}.
\nonumber} \newcommand{\no}{\noindent
\qqq
The ambiguities in this definition are the values
of the integrals
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{_k\over^{4\pi i}}
\int_{_{\widetilde} \def\hat{\widehat B}}\hspace{-0.08cm}\widetilde} \def\hat{\widehat g^*\chi\ -
\ {_k\over^{4\pi i}}\sum\limits_m
\int_{_{S^2_m}}\hspace{-0.1cm}\widetilde} \def\hat{\widehat g|_{_{S^2_m}}^{\, *}
\omega_{_{\lambda_m}}
\label{WZbca}
\qqq
for 3-manifolds $\widetilde} \def\hat{\widehat B$ with $\partial} \newcommand{\ee}{{\rm e}\widetilde} \def\hat{\widehat B=
\mathop{\cup}\limits_m S^2_m$ and for maps
$\widetilde} \def\hat{\widehat g:\widetilde} \def\hat{\widehat B\rightarrow
G$ such that $\widetilde} \def\hat{\widehat g(S^2_m)\subset\CC_{\lambda_m}$,
see Fig.\,\,12.
\leavevmode\epsffile[-7 -20 377 185]{fig12.eps}
\noindent In other words, they are proportional to the periods
of $(\chi,\omega)$ over the cycles of the relative
integer homology $H_3(G,\mathop{\cup}\limits_m
\CC_{\lambda_m})$, as noticed in \cite{KlimSev}.
\vskip 0.3cm
It is not difficult to get a hold on these
ambiguities. Let us glue the unit 3-balls $B_m$
to $\widetilde} \def\hat{\widehat B$ along the boundary spheres $S^2_m$ to obtain
a 3-manifold $\widetilde} \def\hat{\widehat B'$ without boundary and let us extend
$\widetilde} \def\hat{\widehat g$ to a map $\widetilde} \def\hat{\widehat g':\widetilde} \def\hat{\widehat B'\rightarrow G$.
The expression (\ref{WZbca}) may be now rewritten as
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{_k\over^{4\pi i}}
\int_{_{\widetilde} \def\hat{\widehat B'}}\hspace{-0.08cm}\widetilde} \def\hat{\widehat g'{}^*\chi\ -
\ {_k\over^{4\pi i}}\sum\limits_m\Big(\int_{_{B_m}}
\hspace{-0.09cm}{\widetilde} \def\hat{\widehat g'_m}{}^*\chi\ -\,
\int_{_{\partial} \newcommand{\ee}{{\rm e} B_m}}\hspace{-0.09cm}{\widetilde} \def\hat{\widehat g'_m}{}^*
\omega_{\lambda_m}\Big),
\nonumber} \newcommand{\no}{\noindent
\qqq
where $\,\widetilde} \def\hat{\widehat g_m'\,$ are the restrictions of $\widetilde} \def\hat{\widehat g'$
to $B_m$ and they satisfy $\,\widetilde} \def\hat{\widehat g'_m(\partial} \newcommand{\ee}{{\rm e} B_m)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\subset\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
\CC_{\lambda_m}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$. \,As we have discussed in Sect.\,\,3.3,
the first term, involving the integral over the 3-manifold
without boundary $\widetilde} \def\hat{\widehat B'$, takes values in $2\pi i\NZ$
as long as $k$ is an integer.
\vskip 0.3cm
Let us consider the other terms. For $G=SU(2)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}=\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
\{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} x_0+i x_i\sigma_i\,\vert\, x_0^2+x_i^2=1\}\cong S^3$,
the conjugacy classes corresponding to $\,\lambda=j\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
\sigma_3\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$, with $\,0\leq{2j}\leq k\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$, are the 2-spheres
with $\,x_0=\cos{2\pi j\over k}\,$ fixed (except for $j=0$ or
${k\over 2}$ corresponding to the center elements). They
are boundaries of two 3-balls $B_j$
and $B'_j$ with $x_0\geq\cos{2\pi j\over k}$
and $x_0\leq\cos{2\pi j\over k}$, see Fig.\,\,13.
\leavevmode\epsffile[-132 -20 260 215]{fig13.eps}
\noindent A direct calculation shows that
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{_k\over^{4\pi i}}\Big(\int_{_{B_j}}\hspace{-0.1cm}
\chi\,-\,\int_{_{\partial} \newcommand{\ee}{{\rm e} B_j}}\hspace{-0.1cm}\omega\,
\Big)\ =\ -\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} 4\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} j\,.
\qqq
If we used $B'_j$ instead of $B_j$, the result would be
$\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} 4\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}({k\over2}-j)$. We infer that $j$, between $0$
and ${k\over 2}$, must be an integer or a half-integer for
the ambiguity to belong to $2\pi i\NZ$.
This result has already been stated in \cite{AlSch}.
\vskip 0.3cm
For the other groups, the restrictions come from the
the 2-spheres in $\CC_\lambda$ of the form
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\{\, g_0\,\ee^{\m2\pi i\lambda/k}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g_0^{-1}\,\,
\vert\,\,g_0\in SU(2)_\alpha\,\}\,,
\qqq
where $\,SU(2)_\alpha$ is the $SU(2)$ subgroup
of $G$ corresponding to a root $\alpha$, see Sect.\,\,3.3.
Decomposing $\,\lambda=(\lambda-{_1\over^2}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\alpha^\vee\,\tr\,(\alpha
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda))+{_1\over^2}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\alpha^\vee\,\tr\,(\alpha\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$,
\,we observe that the first term commutes
with the generators $\alpha^\vee,\,e_{\pm\alpha}$ of the Lie
algebra of $\,SU(2)_\alpha$ and plays a spectator role.
The calculation of the ambiguity is now essentially
the same as for $G=SU(2)$ with $j$ replaced by
${_1\over^2}\,\tr\,(\alpha\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda)$ and an overall factor
$2\over\hbox{tr}\,\alpha^2$ due to the different
normalization of $\tr\,\alpha^2$. We infer the condition
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{_2\over^{\hbox{tr}\,\alpha^2}}\,\tr\,(\alpha\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda)\ =\
\,\tr\,(\alpha^\vee \lambda)\ \in\ \NZ\,.
\nonumber} \newcommand{\no}{\noindent
\qqq
Since the conjugacy classes $\CC_\lambda$ are in one to one
correspondence with $\lambda$ in the symplex (\ref{sympl}),
the admissible conjugacy classes are in one to one
correspondence with the HW's $\lambda$
integrable at level $k$, see the definition (\ref{iHW}).
\vskip 0.3cm
The full action $S_{_\Sigma}(g)$ of the boundary WZW model
is still obtained by adding to the WZ
action $S^{{WZ}}$ the $S^\gamma$ term of Eq.\,\,(\ref{sgam}).
The coupling to the gauge field is given again by
Eq.\,\,(\ref{tog}). The behavior of the complete
action under the chiral gauge transformation may be shown
to obey the following BC version of the Eq.\,\,(\ref{trprop}):
\begin{eqnarray}} \newcommand{\de}{\bar\partial
S(g,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A)\ =\ S(h_1g\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} h_2^{-1},\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{}^{h_2}
\hspace{-0.1cm}A^{01}+
{}^{h_1}\hspace{-0.09cm}A^{01})\,+\,S(h_1^{-1}h_2,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A)\,
-\,{_{i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} k}\over^{2\pi}}\int_{_\Sigma}\hspace{-0.07cm}
\tr\,({}^{h_2}\hspace{-0.09cm}A^{10}\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{}^{h_1}
\hspace{-0.1cm}A^{01})\,,
\label{trprobc}
\qqq
provided that $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} g\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ satisfies the BC's (\ref{bc}) and that
$\,h_1|_{_{\partial} \newcommand{\ee}{{\rm e} D_{_m}}}=h_2|_{_{\partial} \newcommand{\ee}{{\rm e} D_{_m}}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$. Note that
under this conditions, the field $h_1g\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} h_2^{-1}$
is constrained on the boundary to the same conjugacy classes
as $g$ and $h_1 h_2^{-1}$ to the trivial one so that
the actions on the right hand side make sense. The above
relation will be employed below to infer the chiral gauge
symmetry Ward identities for the boundary WZW theory. It may
be used as the principle that selects the BC's (\ref{bc}).
\vskip 0.3cm
Summarizing: if the field $g:\Sigma\rightarrow G$ satisfies
the BC's (\ref{bc}) with integrable weights $\lambda_m$,
then the amplitude $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\ee^{-\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} S^{^{WZ}}_{_\Sigma}(g)}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$,
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} and consequently also $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\ee^{-\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} S_{_\Sigma}(g)}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ and
$\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\ee^{-\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} S_{_\Sigma}(g,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A)}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$,
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} may be well defined as complex numbers. Of course,
the mixed case, where the BC's (\ref{bc}) with integrable
$\lambda_m$ are satisfied only on some boundary components
of $\Sigma$ and no conditions are imposed on the other
(``free'') boundary components can be treated in the same
way. It results in the amplitudes
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\ee^{-\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} S_{_\Sigma}(g,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A)}\ \in\
\mathop{\otimes}
\limits_{n\ {\rm free}}
{\cal L}} \newcommand{\CM}{{\cal M}_{_{g|_{_{\partial} \newcommand{\ee}{{\rm e} D_{_n}}}}}\,.
\nonumber} \newcommand{\no}{\noindent
\qqq
\vskip 0.4cm
\subsection{Quantum amplitudes and correlation functions}
The functional integral definition (\ref{FfiA}) of the quantum
amplitudes of the WZW model may be naturally
generalized to the case where on some boundary components
of $\Sigma$ we impose the BC's (\ref{bc}) with integrable
weights $\lambda_m$. The resulting amplitudes
$\,\CA_{_{\Sigma,\un{\lambda}}}\hspace{-0.09cm}
(A)\,$ will be now elements
of $\mathop{\otimes}\limits_{n\ {\rm free}}\Gamma({\cal L}} \newcommand{\CM}{{\cal M})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$.
\,They may be represented as (partial) contractions of the
amplitudes $\,\CA_{_\Sigma}(A)\,$
with all boundaries free with appropriate states:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\CA_{_{\Sigma,\un{\lambda}}}\hspace{-0.09cm}(A)\ =\
(\,\mathop{\otimes}\limits_m \hat\delta_{_{\lambda_m}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},\,
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\CA_{_\Sigma}(A)\,)\,.
\nonumber} \newcommand{\no}{\noindent
\qqq
The non-normalizable states\footnote{technically, antilinear
forms on a dense subspace of the Hilbert space ${\cal H}} \newcommand{\CI}{{\cal I}$}
$\hat\delta_{_{\lambda}}$ are given
by Cardy's formula \cite{Cardy}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\hat\delta_{_\lambda}\ \ =\ \
\sum\limits_{\hat R}\,(S^{^{1}}_{_{R}})^{-{_1\over^2}}\,
S^{^{R_\lambda}}_{_{R}}
\,\,e^{\hat{i}}_{_{\hat R}}
\otimes\overline{e^{\hat{i}}_{_{\hat R}}}\,,
\label{Card}
\qqq
where $R_\lambda$ denotes the representation
of $G$ with the HW $\lambda$, the vectors
$\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} e^{\hat{i}}_{_{\hat R}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ form an orthonormal
basis of the space $V_{_{\hat{R}}}$
and the sum over $\hat{i}$ is understood.
Note the analogy with Eq.\,\,(\ref{Ishcl})
for the delta-function supported by a conjugacy class
$\CC_\lambda$. The matrix $(S^{^{R_\lambda}}_{_{R}})$
has replaced the one with the elements
$\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}(\chi_{_R}(\ee^{\m2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda/k})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ and
the representation spaces $V_{_{\hat{R}}}$ of the current
algebras those of the finite-dimensional group.
The state ${\hat\delta}_{_\lambda}$ should be interpreted
as a delta-function concentrated on the loops
in $LG$ contained in the conjugacy
class $\CC_\lambda$. The non-normalizable states
$\,e^{\hat{i}}_{_{\hat R}}\otimes\overline
{e^{\hat{i}}_{_{\hat R}}}\,$ are called the Ishibashi states
\cite{Ishib} and generalize the (properly normalized)
characters of the group,
see Eq.\,\,(\ref{chdec}).
\vskip 0.3cm
The correlation functions $\,<\otimes g_{_{R_n}}
\hspace{-0.05cm}(\xi_n)>_{_{\hspace{-0.08cm}\Sigma,
\un{\lambda}}}\hspace{-0.19cm}(A)\,$
in the presence of the boundaries with fields constrained
by the the BC's (\ref{bc}) may be again defined
by the functional integral (\ref{cfu}) taking numerical
values. The transformation property (\ref{trprobc})
of the action implies now that
\begin{eqnarray}} \newcommand{\de}{\bar\partial
&&{\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{\Sigma,\un{\lambda}}}
\hspace{-0.09cm}(A)\,
<\mathop{\otimes}\limits_n
g_{_{R_n}}\hspace{-0.05cm}(\xi_n)>_{_{\hspace{-0.08cm}\Sigma,
\un{\lambda}}}\hspace{-0.09cm}(A)\ \,=\ \,
\ee^{-S(h_1^{-1}h_2,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A)
\,+\,{_{i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} k}\over^{2\pi}}\int_{_\Sigma}\hspace{-0.07cm}
\tr\,({}^{^{h_2}}\hspace{-0.1cm}A^{10}\,{}^{^{h_1}}
\hspace{-0.11cm}A^{01})}
\ {\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{\Sigma,\un{\lambda}}}({}^{h_2}\hspace{-0.09cm}A^{10}+
{}^{h_1}\hspace{-0.1cm}A^{01})\hspace{0.3cm}
\cr\cr
&&\hspace{2.4cm}\cdot\ \mathop{\otimes}\limits_n
(h_1)_{_{R_n}}^{-1}
\hspace{-0.03cm}(\xi_n)\ \
<\mathop{\otimes}\limits_n
g_{_{R_n}}\hspace{-0.05cm}(\xi_n)
>_{_{\hspace{-0.08cm}\Sigma,\un{\lambda}}}
\hspace{-0.1cm}({}^{h_2}\hspace{-0.09cm}A^{01}+
{}^{h_1}\hspace{-0.1cm}A^{01})\ \
\mathop{\otimes}\limits_n (h_2)_{_{R_n}}
\hspace{-0.05cm}(\xi_n)\,.
\label{WIbc}
\qqq
This is the chiral gauge symmetry Ward identity
for the boundary WZW correlation functions,
a variant of the identities (\ref{WIcfl}) and (\ref{WIcfr})
in presence of the BC's. Note, however, that
the identity (\ref{WIbc}) may be factorized as
the latter ones only if we assume that $h_1$
and $h_2$ are equal to $1$ on the boundary. The general
case where on the boundary $h_1=h_2$ requires the presence
of both gauge transformations $h_1$ and $h_2$.
\vskip 0.3cm
It is illuminating to rewrite the Ward identity
(\ref{WIbc}) in a different way. To this end,
let us define a ``doubled'' Riemann surface
without boundary $\widetilde} \def\hat{\widehat\Sigma$ by gluing $\Sigma$
to its complex conjugate $\overline{\Sigma}$ along
the boundary components, see Fig.\,\,14.
\leavevmode\epsffile[-102 -20 320 210]{fig14.eps}
We shall denote by $\,\iota\,$
the anti-holomorphic involution of $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\widetilde} \def\hat{\widehat\Sigma$
exchanging $\Sigma$ with its complex conjugate:
$\,\iota(\xi)=\bar\xi\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$. \,Each chiral gauge
field $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\widetilde} \def\hat{\widehat A^{01}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ on the Riemann surface
$\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\widetilde} \def\hat{\widehat\Sigma\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ defines a complexified gauge field
$\,A=(\iota^*A^{01}+A^{01})\vert_{_\Sigma}\,$
on the surface $\Sigma$. Let us define
a holomorphic functional
$$\Psi_{_{\un{\lambda}}}:\,\widetilde} \def\hat{\widehat\CA^{01}\
\longrightarrow\ (\mathop{\otimes}\limits_n
V_{_{R_n}})\otimes(\mathop{\otimes}\limits_n
V_{_{\overline{R}_n}})$$
of the chiral gauge fields on the doubled surface by
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\Psi_{_{\un{\lambda}}}(\widetilde} \def\hat{\widehat A^{01})\ =\
{\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{\Sigma,\un{\lambda}}}
\hspace{-0.09cm}(A)\,<\mathop{\otimes}\limits_n
g_{_{R_n}}\hspace{-0.05cm}(\xi_n)>_{_{\hspace{-0.08cm}\Sigma,
\un{\lambda}}}\hspace{-0.09cm}(A)\,,
\nonumber} \newcommand{\no}{\noindent
\qqq
where we identify the space $\,(\otimes
V_{_{R_n}})\otimes(\otimes V_{_{\overline{R}_n}})\,$
with $\,\otimes End(V_{_{R_n}})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$. \,Let $\widetilde} \def\hat{\widehat h$
be a $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} G^\NC$-valued gauge transformation on $\widetilde} \def\hat{\widehat\Sigma$.
We shall pose $h_1=\widetilde} \def\hat{\widehat h\vert_{_\Sigma}$
and $h_2=\iota^*\widetilde} \def\hat{\widehat h\vert_{_\Sigma}$. Note that $h_1=h_2$
on the boundary of $\Sigma$. It is not difficult
to prove that
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\ee^{-S_{_\Sigma}(h_1^{-1}h_2,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A)
\,+\,{_{i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} k}\over^{2\pi}}\int_{_\Sigma}\hspace{-0.07cm}
\tr\,({}^{^{h_2}}\hspace{-0.1cm}A^{10}\,{}^{^{h_1}}
\hspace{-0.11cm}A^{01})}\ =\ \ee^{-S_{_{\widetilde} \def\hat{\widehat\Sigma}}
(\widetilde} \def\hat{\widehat h^{-1},\,\widetilde} \def\hat{\widehat A^{01})}\,.
\qqq
The Ward identity (\ref{WIbc}) implies then that
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\Psi_{_{\un{\lambda}}}(\widetilde} \def\hat{\widehat A^{01})\ =\
\ =\ \ee^{-S_{_{\widetilde} \def\hat{\widehat\Sigma}}
(\widetilde} \def\hat{\widehat h^{-1},\,{\widetilde} \def\hat{\widehat A}^{01})}\
(\mathop{\otimes}\limits_n \widetilde} \def\hat{\widehat h^{-1}_{_{R_n}}
(\xi_n)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch})
\otimes(\mathop{\otimes}\limits_n
\widetilde} \def\hat{\widehat h^{-1}_{_{\overline{R}_n}}(\bar\xi_n)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch})\
\Psi_{_{\un{\lambda}}}({}^{\widetilde} \def\hat{\widehat h}\hspace{-0.09cm}
{\widetilde} \def\hat{\widehat A}^{01})\,,
\nonumber} \newcommand{\no}{\noindent
\qqq
i.e.\,\, that $\,\Psi_{_{\un{\lambda}}}\,$ is a CS state
on the doubled surface $\widetilde} \def\hat{\widehat\Sigma$ with the doubled
insertions at points $\xi_n$ and at their complex conjugates
$\bar\xi_n$, associated, respectively,
to the representations $R_n$ and to the complex conjugate
representations $\overline{R}_n$.
\vskip 0.3cm
We infer that the correlations functions of the boundary
WZW theory on a surface $\Sigma$ may be viewed,
in their gauge field dependence, as the special states
$\,\Psi_{_{\un{\lambda}}}\,$ belonging to the space
$\,\NW_{_{\widetilde} \def\hat{\widehat\Sigma}}(\un{\xi},\un{\bar\xi},\un{R},\un{\overline{R}})\,$
of the CS states on the doubled surface $\widetilde} \def\hat{\widehat\Sigma$. The states
$\,\Psi_{_{\un{\lambda}}}\,$ may be shown, moreover, to be preserved
by the parallel transport with respect to the KZB connection on the
restriction of the bundle $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\widetilde} \def\hat{\widehat\CW\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$
of the CS states on the doubled surface to the subspace
of the ``doubled'' complex structures and insertions.
These properties are often summarized by saying
that the boundary CFT is chiral since its
correlation functions are given by special conformal blocks
of the WZW theory on $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\widetilde} \def\hat{\widehat\Sigma$. It would
be desirable to characterize geometrically the CS states
$\,\Psi_{_{\un{\lambda}}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$. \,Some of their
special properties are easy to find. For example,
they are preserved by the antilinear involution
$\,\Psi\mapsto{}^\iota\Psi\,$ of $\,\NW_{_{\widetilde} \def\hat{\widehat\Sigma}}(\un{\xi},
\un{\bar\xi},\un{R},\un{\overline{R}})\,$ induced by the involution
$\,\iota\,$ of the doubled surface $\widetilde} \def\hat{\widehat\Sigma$ and defined by
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{}^\iota\Psi(\widetilde} \def\hat{\widehat A^{01})\ =\ \overline{\Psi(-(\iota^*\widetilde} \def\hat{\widehat
A^{01})^*)}\,.
\qqq
A complete geometric characterization of the states
$\,\Psi_{_{\un{\lambda}}}\,$ for different choices $\un{\lambda}$
of the BC's seems, however, still missing.
\subsection{Piece-wise boundary conditions}
Up to now, we have imposed the boundary conditions
forcing the fields to take values in the special conjugacy
classes uniformly on the component circles of $\partial} \newcommand{\ee}{{\rm e}\Sigma$.
Since the conditions are local, it should be also possible
to do this locally on the pieces of the boundary. Suppose
that the boundary $\partial} \newcommand{\ee}{{\rm e}\Sigma$ is divided into
intervals $I_r$ (the entire boundary circles are
also admitted). We shall associate integrable weight
labels $\lambda_r$ to some of these intervals in such a way
that two labeled intervals in the same boundary component
are separated by an unlabeled (``free'') one,
see Fig.\,\,15.
\leavevmode\epsffile[-80 -20 337 210]{fig15.eps}
\noindent We shall now consider the fields
$g$ on $\Sigma$ which on the labeled intervals
take values in the corresponding conjugacy classes
$\CC_{\lambda_r}$ and are not restricted on the free
intervals. One may still define the classical amplitudes
$\,\ee^{-S^{{WZ}}(g)}\,$ for such fields although this
requires a more local ($\check{{\rm C}}$ech cohomology type)
technique than the one developed above
\cite{Carg}. Let us sketch how this is done.
\vskip 0.1cm
\leavevmode\epsffile[-37 -20 357 220]{fig16.eps}
Recall from the end of Sect.\,\,3.1 that $S^{^{WZ}}(g)$ in the
first approximation is equal to $\,{k\over 4\pi i}\int g^*\beta$
where $\,d\beta=\chi\,$ is the canonical closed 3-form on $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} G$.
The problem stemmed from the fact that such 2-forms $\beta$
exist only locally. However, on the sets of a sufficiently
fine open covering $(\CO_\alpha)$ of $G$, we
may choose 2-forms \s$\beta_\alpha\s$ such that
$d\beta_\alpha=\chi$. Choose a triangulation ${\cal T}} \newcommand{\CU}{{\cal U}$ of $\Sigma$
with the triangles $t$, edges $e$ and vertices $v$. If
${\cal T}} \newcommand{\CU}{{\cal U}$ is fine enough then each of the simplices $s$ of ${\cal T}} \newcommand{\CU}{{\cal U}$
is mapped by $g$ into an open set, say $\CO_{\alpha_s}$,
see Fig.\,\,16. The main contribution to the amplitude
$\,\ee^{-S^{^{WZ}}_{_\Sigma}(g)}\,$
will come from $\,\exp[-{k\over4\pi i}\sum_{_t}
\hspace{-0.06cm}\int_{_t}g^*\beta_{\alpha_t}]\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}.$
\,The above expression depends, however, on the choice of
the forms $\,\beta_\alpha\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ and of the triangulation. The
idea is to compensate this dependence by contributions from
simplices of lower dimension. Let
\s$\eta_{\alpha_{_0}\alpha_{_1}}=
-\eta_{\alpha_{_1}\alpha_{_0}}\s$ be 1-forms defined on
the non-empty
intersections \s$\CO_{\alpha_{_0}\alpha_{_1}}
\equiv\CO_{\alpha_{_0}}\cap\CO_{\alpha_{_1}}\s$ such that
$$d\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\eta_{\alpha_{_0}\alpha_{_1}}\ =\ \beta_{\alpha_{_1}}-
\beta_{\alpha_{_0}}$$
and let $\,f_{\alpha_{_0}\alpha_{_1}\alpha_{_2}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$
be functions on the triple
intersections $\,\CO_{\alpha_{_0}\alpha_{_1}\alpha_{_2}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$,
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} antisymmetric in the indices, satisfying
$$d\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} f_{\alpha_{_0}\alpha_{_1}\alpha_{_2}}\ =\
\eta_{\alpha_{_1}\alpha_{_2}}-\eta_{\alpha_{_0}
\alpha_{_2}}+\eta_{\alpha_{_0}\alpha_{_1}}$$
and such that on the four-fold intersections
$$f_{\alpha_{_1}\alpha_{_2}\alpha_{_3}}-
f_{\alpha_{_0}\alpha_{_2}\alpha_{_3}}+f_{\alpha_{_0}
\alpha_{_1}\alpha_{_3}}-
f_{\alpha_{_0}\alpha_{_1}\alpha_{_2}}\in 8\pi^2\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\NZ\,.$$
Such data may, indeed, be chosen.
We define then
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\ee^{-S_{_\Sigma}^{^{WZ}}(g)}\,=\,
\exp\Big[-{_k\over^{4\pi i}}\Big(\sum\limits_t\hspace{-0.06cm}
\int\limits_{t}\hspace{-0.06cm}g^*\beta_{\alpha_t}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
-\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\sum\limits_{e\subset t}
\int\limits_e\hspace{-0.06cm}g^*\eta_{\alpha_e\alpha_t}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}+\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
\sum\limits_{v\in e\subset t}\hspace{-0.06cm}(\pm)
f_{\alpha_v\alpha_e\alpha_t}(g(v))\Big)\Big]\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},\hspace{0.5cm}
\label{WZ2}
\qqq
where in the last sum the sign is taken according to the
orientation of the vertices $v$ inherited from the triangles $t$
via the edges $e$. One may show that for the surface without
boundary, the above expression does not depend on the choices
involved and coincides with the definition given in Sect.\,\,3.3.
\vskip 0.3cm
In the presence
of boundary circles with unconstrained fields, the above
expression may be used to define the amplitudes with
values in a line bundle and it provides an alternative
construction of the bundle ${\cal L}} \newcommand{\CM}{{\cal M}$ over the loop group \cite{Carg}.
In the presence of the boundary conditions on the intervals
$I_r$ we shall still employ the same definition, but with
some care. Namely, we include neighborhoods $\CO_\lambda$
of the conjugacy classes $\CC_\lambda$ into the open covering
$(\CO_\alpha)$ of $G$. We choose 2-forms $\beta_\lambda$
on $\CO_\lambda$ coinciding with $\omega_\lambda$
of Eq.\,\,(\ref{om}) when restricted to $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\CC_\lambda$.
The triangulations used in Eq.\,\,(\ref{WZ2}) are
required to be compatible with the splitting of the boundary.
To the simplices in the labeled
intervals $I_r$ we asign the open sets $\CO_{\lambda_r}$,
see Fig.\,\,16.
The amplitudes resulting from Eq.\,\,(\ref{WZ2}) coincide
then with those defined in the previous
section for the special case when the labeled intervals
fill entire circles. In the general case,
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\ee^{-S^{^{WZ}}_{_\Sigma}(g)}\ \ \in\ \
\prod\limits_{{\rm free}\ I_r}
({\cal L}} \newcommand{\CM}{{\cal M}_{_{I_r}}
)_{_{g\vert_{_{I_r}}}}
\qqq
where ${\cal L}} \newcommand{\CM}{{\cal M}_{_{I_r}}$ is a line bundle
over the space of maps from an interval $I_r$ to $G$
taking on the boundary of $I_r$ the values
in the conjugacy classes $\CC_{\lambda_r}$ and
$\CC_{\lambda'_r}$
specified by the labels of the neighboring
intervals\footnote{if $I_r$ is a full circle,
${\cal L}} \newcommand{\CM}{{\cal M}_{_{I_r}}={\cal L}} \newcommand{\CM}{{\cal M}$}.
\vskip 0.3cm
The space of sections $\Gamma({\cal L}} \newcommand{\CM}{{\cal M}_{_{I_r}})$ plays,
as before, the role of the space of states of the WZW theory
but, this time, on the interval and with boundary conditions
specified by the conjugacy classes $\CC_{\lambda_r}$
and $\CC_{\lambda'_r}$.
In the string language, these are states of the open string
moving on the group with the ends on the {\bf branes}
$\CC_{\lambda_r}$ and $\CC_{\lambda'_r}$, see Fig.\,\,17.
\leavevmode\epsffile[-70 -20 310 156]{fig17.eps}
\noindent One may still define an action of the central extension
of the loop group in the spaces $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Gamma({\cal L}} \newcommand{\CM}{{\cal M}_{_{I_r}})\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$
(a single one) and base on its analysis
a rigorous construction of the open-string Hilbert
spaces of states ${\cal H}} \newcommand{\CI}{{\cal I}_{_{\lambda\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda'}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$,
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} as we did in Sect.\,\,4.1 for the closed-string
states, see Eq.\,\,(\ref{HSrd}). One obtains
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{\cal H}} \newcommand{\CI}{{\cal I}_{_{\lambda\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda'}}\ =\
\mathop{\oplus}\limits_{_{\hat R}}\,
M_{\lambda\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda'}^{^{\,R}}\otimes V_{_{\hat R}}\,.
\label{opHs}
\qqq
The multiplicity spaces may be naturally identified
with the spaces of the genus zero CS states
$\CW(\NC P^1,\un{\xi},\un{R})$ with three insertion
points in representations $\overline{R}_{\lambda}$, $R_{\lambda'}$
and $R$. In particular, the dimension of the multiplicity
spaces is given by the fusion ring structure constants
$\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\hat N}^{\,\overline{R}}_{\overline{R}_{\lambda}R_{\lambda'}}$.
The spaces ${\cal H}} \newcommand{\CI}{{\cal I}_{_{\lambda\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\lambda'}}$
carry the obvious action of the current algebra $\hat{{\bf g}}} \newcommand{\Ntt}{{{\bf t}}$
and of the Virasoro algebra, the latter obtained
by the Sugawara construction (\ref{sug}). The generator
$L_0-{c\over 24}$ gives the Hamiltonian of the open string
sectors. The spaces ${\cal H}} \newcommand{\CI}{{\cal I}_{\lambda\lambda}$ with the same
BC on both sides contain the vacua $\Omega_{_\lambda}$,
i.e.\,\,the states annihilated by $L_0$ (unique up
to normalization).
\vskip 0.4cm
\subsection{Elementary quantum amplitudes}
The quantum amplitudes with the general boundary conditions
are given now by the formal functional integrals.
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\CA_{_{\Sigma,\un{I},\un{\lambda}}}\hspace{-0.09cm}(A)\ \ =\
\int\limits_{g(I_r)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\subset\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\CC_{\lambda_r}}
\ee^{-S_{_\Sigma}(g,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A)}\,\, Dg
\qqq
and, should take values in the space
$\,\mathop{\otimes}
\limits_{{\rm free}\ I_r}{\cal H}} \newcommand{\CI}{{\cal I}_{_{\lambda_r\lambda'_r}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$.
\,They should possess a gluing property
along free boundary intervals with
opposite boundary weight assignment, generalizing
the gluing properties (\ref{glueB}) or (\ref{glueC}).
As discussed in detail by Segal in \cite{Segal}
for the closed string sector,
the general amplitudes may be constructed by gluing
from the elementary ones for the geometries listed
on Fig.\,\,18.
\leavevmode\epsffile[-40 -10 345 290]{fig18.eps}
\vskip 0.1cm
The elementary amplitudes (a) and (b) represent, respectively,
the vacuum state $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Omega\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ in the closed string space ${\cal H}} \newcommand{\CI}{{\cal I}$,
and the vacua $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Omega_{_\lambda}$ in the open string
spaces ${\cal H}} \newcommand{\CI}{{\cal I}_{_{\lambda\lambda}}$.
The amplitudes (c) for arbitrary annuli encode the
action of the pair of Virasoro algebras in ${\cal H}} \newcommand{\CI}{{\cal I}$.
In particular, for a complex number $q\not=0$ inside
the unit disc one may consider the annular regions
$\,A_q=\{\,z\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\vert\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\,
\vert q\vert\leq\vert z\vert\leq1\,\}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$, \hspace{0.025cm}} \newcommand{\ch}{{\rm ch} see Fig.\,\,19,
obtained from $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\NC P^1$ by taking out the unit discs
embedded by the maps $z\mapsto qz$ and $z\mapsto z^{-1}$.
\leavevmode\epsffile[-53 -20 310 165]{fig19.eps}
\noindent Viewing the amplitude of $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A_q$ as
an operator from the space ${\cal H}} \newcommand{\CI}{{\cal I}$ associated to the first
boundary component to ${\cal H}} \newcommand{\CI}{{\cal I}$ associated to the second
boundary, one has:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{A_q}}^{-1}\,\,\CA_{_{A_q}}\ =
\ q^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} L_0}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\bar q}^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\widetilde} \def\hat{\widehat L_0}\,.
\nonumber} \newcommand{\no}{\noindent
\qqq
The gluing of the two boundary circles of $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} A_q$ leads
to the complex torus $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{T}_\tau$ where
$q=\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} 2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\tau}$.
According to the gluing relation (\ref{glueC}), this
produces the toroidal partition function
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{\cal Z}} \newcommand{\s}{\hspace{0.05cm}({T}_\tau)\ =\ {\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{A_q}}\ \tr_{_{\cal H}} \newcommand{\CI}{{\cal I}}\,(q^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} L_0}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
{\bar q}^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\widetilde} \def\hat{\widehat L_0})\,.
\qqq
Upon choosing a flat metric on ${T}_\tau$ and
working out the partition function\footnote{${\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{A_q}}$
is a ratio of two partition function on the Riemann sphere
and may be easily found from the relation (\ref{scin})
to be equal to $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\vert q\vert^{-{c\over12}}$}
${\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{A_q}}$, one obtains finally
\begin{eqnarray}} \newcommand{\de}{\bar\partial
{\cal Z}} \newcommand{\s}{\hspace{0.05cm}(\tau)\ =\ \tr_{_{{\cal H}} \newcommand{\CI}{{\cal I}}}\,q^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} L_0-{c\over 24}}\,
{\bar q}^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\widetilde} \def\hat{\widehat L_0-{c\over 24}}
\qqq
which is nothing else but Eq.\,\,(\ref{diag}).
\vskip 0.3cm
The amplitude for a disc $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} P_{w,q,q_1}$ with two round
holes, as in Fig.\,\,18(d), gives rise to a 3-linear form
on ${\cal H}} \newcommand{\CI}{{\cal I}$ which may be also viewed as an operator from
the space ${\cal H}} \newcommand{\CI}{{\cal I}\otimes{\cal H}} \newcommand{\CI}{{\cal I}$ associated to the inner discs
to ${\cal H}} \newcommand{\CI}{{\cal I}$ corresponding to the outer one. It is customary
in CFT to rewrite this amplitude as an operator $\Phi(e;\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} w)$
in ${\cal H}} \newcommand{\CI}{{\cal I}$ labeled by the vectors $e$ in (a dense subspace of)
${\cal H}} \newcommand{\CI}{{\cal I}$ and the point $w$ inside the unit disc:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\Phi\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}(e;\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} w)\, e'\
=\ {\cal Z}} \newcommand{\s}{\hspace{0.05cm}_{_{P_{w,q,q_1}}}\,\,\CA_{_{P_{w,q,q_1}}}
\,(q^{-L_0}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\bar q}^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}-\widetilde} \def\hat{\widehat L_0}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} e')\otimes
(q_1^{-L_0}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\bar q_1}^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}-\widetilde} \def\hat{\widehat L_0}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} e)\,.
\qqq
The combination with the powers of $L_0$ and $\widetilde} \def\hat{\widehat L_0$
assures the independence of the expression of $q$ and $q_1$.
The vectors $e$ can be recovered from the operators
$\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Phi(e;\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} w)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ by acting with them on the vacuum
vector
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\lim\limits_{w\to0}\,\Phi(e;\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} w)\,\Omega\ =\ e\,.
\qqq
Pictorially, this corresponds to filling up the
central whole of $P_{w,q,q_1}$ by gluing a disc to its
boundary. The operators $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Phi(e;\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} w)\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ satisfy an important
relation:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\Phi\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}(e;\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} z)\,\,\Phi(e';\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} w)
\ =\ \Phi(\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Phi(e;\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} z-w)\,e';\,w)\,.
\label{OPEgl}
\qqq
The above identity holds for $\m0<|w|<|z|\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$ and $\m0<|z-w|<1$.
It results from the two ways that one may obtain the disc with
three holes by gluing two discs with two holes, see Fig.\,\,20.
\leavevmode\epsffile[-65 -20 320 160]{fig20.eps}
\noindent The relation (\ref{OPEgl}) may be viewed as a
global form of the operator product expansion.
The local forms may be extracted from it by expanding
the vector $\,\Phi(e;\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} z-w)\,$ into terms homogeneous
in $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}(z-w)$. \hspace{0.025cm}} \newcommand{\ch}{{\rm ch} In particular,
for specially chosen vectors $e$ and $e'$ one obtains
the operator versions of the relations
(\ref{fope}) and (\ref{opem}), hence the name of the latter.
The vector-operator correspondence together with the
operator product expansion (\ref{OPEgl})
are the cornerstones of the non-perturbative
approach to CFT.
\vskip 0.3cm
The amplitudes corresponding to the surfaces with
boundary of Fig.\,\,18(e) represent the action
of the Virasoro algebra in the open string spaces
$\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\cal H}} \newcommand{\CI}{{\cal I}_{_{\lambda\lambda'}}$.
The surfaces (f) give rise, in turn, to the amplitudes
which, applied to vectors $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} q_1^{-L_0}
{\bar q_1}^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}-\widetilde} \def\hat{\widehat L_0}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} e\in{\cal H}} \newcommand{\CI}{{\cal I}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}$, \,define
the action of the closed string sector fields
$\,\Phi(e,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} w)\,$ in the open string
spaces $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\cal H}} \newcommand{\CI}{{\cal I}_{_{\lambda\lambda'}}$. Finally,
the amplitudes of the disc (g) with three
labeled and three free boundary intervals
define 3-linear forms on the corresponding
open string spaces. As before in the closed
string sector, one may interpret them in terms
of boundary operators labeled by vectors
in, say, $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\cal H}} \newcommand{\CI}{{\cal I}_{_{\lambda''\lambda}}$
and mapping from $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\cal H}} \newcommand{\CI}{{\cal I}_{_{\lambda\lambda'}}$
to $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\cal H}} \newcommand{\CI}{{\cal I}_{_{\lambda'\lambda''}}$.
\vskip 0.3cm
The gluing properties give rise to non-trivial
relations between various amplitudes. For example,
gluing along the free sides a rectangle
with a local BC imposed on the two other sides,
see Fig.\,\,21, one obtain a finite cylinder $Z_{_L}$
with the BC's imposed on the boundary components.
\leavevmode\epsffile[-44 -20 330 140]{fig21.eps}
\noindent Its amplitude $\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\CA_{_{Z_{_{L}}}}$ (in
the flat metric) may be computed in two ways. On one
hand side, using the decomposition (\ref{opHs}),
we infer that
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\CA_{_{Z_{_{L}}}}\ =\ \tr_{_{H_{_{\lambda\lambda'}}}}\,
q^{L_0-{c\over 24}}\ =\ \sum\limits_{\hat{R}}\,
{\hat N}^{\,\overline{R}}_{\overline{R}_{\lambda}R_{\lambda'}}\,\,
\tr_{_{V_{_{\hat{R}}}}}\,q^{L_0-{c\over 24}}
\ =\ \sum\limits_{\hat{R}}\,
{\hat N}^{\,\overline{R}}_{\overline{R}_{\lambda}R_{\lambda'}}\,\,
\chi_{_{\hat{R}}}(\tau,\m1)\hspace{0.4cm}
\label{Car1}
\qqq
with $\tau={L\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} i\over2\pi}$ and $q=\ee^{\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} 2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\tau}$.
On the other hand, we may express this amplitude
as a matrix element of the close string amplitude
between the boundary states $\hat{\delta}_{_\lambda}$ and
$\hat{\delta}_{_{\lambda'}}$. With $\,q'=\ee^{\m2\pi i\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\tau'}$
and $\,\tau'=-{1\over\tau}={2\pi i\over L}$, \,we obtain
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\CA_{_{Z_{_{L}}}}\ =\ \Big(\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\hat{\delta}_{_{\lambda}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},
\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}(q')^{{_1\over^2}(L_0-{c\over24})}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
(\overline{q'})^{{_1\over^2}(\widetilde} \def\hat{\widehat L_0-{c\over24})}\,
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\hat{\delta}_{_{\lambda'}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Big)\,.
\qqq
Upon the substitution of Cardy's expression (\ref{Card})
for the boundary states $\hat{\delta}_{_\lambda}$, this
becomes
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\CA_{_{Z_{_{L}}}}
&=&
\sum\limits_{\hat{R},\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\hat{R}}'}
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}(S^{^{1}}_{_{R}})^{-{_1\over^2}}\,
S^{^{\overline{R}_\lambda}}_{_{R}}
\,(S^{^{1}}_{_{R'}})^{-{_1\over^2}}\,
S^{^{R_{\lambda'}}}_{_{R'}}
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\,\Big(\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} e^{\hat{i}}_{_{\hat R}}
\otimes\overline{e^{\hat{i}}_{_{\hat R}}}\,,
\,\,(q')^{{_1\over^2}(L_0-{c\over24})}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}
(\overline{q'})^{{_1\over^2}(\widetilde} \def\hat{\widehat L_0-{c\over24})}\,
e^{{\hat{i}}'}_{_{{\hat{R}}'}}
\otimes\overline{e^{{\hat{i}}'}_{_{{\hat{R}}'}}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Big)
\hspace{0.4cm}\cr
&=&
\sum\limits_{{\hat{R}}'}
\,(S^{^{1}}_{_{R'}})^{-1}\,
S^{^{\overline{R}_\lambda}}_{_{R'}}
\,S^{^{R_{\lambda'}}}_{_{R'}}
\,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\Big(e^{\hat{i}}_{_{\hat R'}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},
\,(q')^{{_1\over^2}(L_0-{c\over24})}
\,e^{{\hat{i}}'}_{_{{\hat{R}}'}}\Big)\,
\Big(e^{{\hat{i}}'}_{_{{\hat{R}}'}}\hspace{0.025cm}} \newcommand{\ch}{{\rm ch},
\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}\,(\overline{{q'}})^{{_1\over^2}(L_0-{c\over24})}
\,e^{\hat{i}}_{_{{\hat R}'}}\Big)\cr
&=&
\sum\limits_{{\hat{R}}'}
\,(S^{^{1}}_{_{R'}})^{-1}\,
S^{^{\overline{R}_\lambda}}_{_{R'}}
\,S^{^{R_{\lambda'}}}_{_{R'}}
\,\,\tr_{_{V_{_{{\hat R}'}}}}\,(q')^{L_0-{c\over24}}
\ =\
\sum\limits_{{\hat{R}}'}
\,(S^{^{1}}_{_{R'}})^{-1}\,
S^{^{\overline{R}_\lambda}}_{_{R'}}
\,S^{^{R_{\lambda'}}}_{_{R'}}
\,\,\chi_{_{{\hat R}'}}(\tau',\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} 1)\,.
\qqq
With the use of the modular transformation property
(\ref{modm}), we finally obtain:
\begin{eqnarray}} \newcommand{\de}{\bar\partial
\CA_{_{Z_{_{L}}}}\ =\
\sum\limits_{\hat{R},\hspace{0.025cm}} \newcommand{\ch}{{\rm ch}{\hat R}'}
\,(S^{^{1}}_{_{R'}})^{-1}\,
S^{^{\overline{R}_\lambda}}_{_{R'}}
\,S^{^{R_{\lambda'}}}_{_{R'}}
\,S^{^{R}}_{_{R'}}\,\,\chi_{_{\hat R}}(\tau,\hspace{0.025cm}} \newcommand{\ch}{{\rm ch} 1)\,.
\nonumber} \newcommand{\no}{\noindent
\qqq
By virtue of the Verlinde formula (\ref{orV}),
the last identity coincides with Eq.\,\,(\ref{Car1}).
We have, in fact, inverted here the logic of reference
\cite{Cardy}, where the consistency of the two ways
of computing the amplitude $\,\CA_{_{Z_{_{L}}}}$ was used
to obtain the expression (\ref{Card}) for the boundary
states $\,\hat{\delta}_{_\lambda}$.
\vskip 0.3cm
The whole system of elementary amplitudes represents
an intriguing algebraic structure which is
common to all (rational) boundary CFT's. Already the case
of boundary topological field theories, where the amplitudes
depend only on the surface topology, leads to an interesting
construction that remains to be fully understood. It
entangles a commutative algebra structure on the closed
string space of states and a non-commutative algebroid
in the open string sector. An example of such a structure
was inherent in the work of Kontsevich \cite{Konts2}
on the deformation quantization of general
Poisson manifolds, see \cite{Felder}. Certainly,
the two-dimensional CFT did not unveal yet all of its
secrets.
\vskip 1cm
|
\section{#2}\label{S:#1}%
\ifShowLabels \TeXref{{S:#1}} \fi}
\newcommand{\ssec}[2]{\subsection{#2}\label{SS:#1}%
\ifShowLabels \TeXref{{SS:#1}} \fi}
\newcommand{\refs}[1]{Section ~\ref{S:#1}}
\newcommand{\refss}[1]{Subsection ~\ref{SS:#1}}
\newcommand{\reft}[1]{Theorem ~\ref{T:#1}}
\newcommand{\refl}[1]{Lemma ~\ref{L:#1}}
\newcommand{\refp}[1]{Proposition ~\ref{P:#1}}
\newcommand{\refc}[1]{Corollary ~\ref{C:#1}}
\newcommand{\refconj}[1]{Conjecture ~\ref{Conj:#1}}
\newcommand{\refd}[1]{Definition ~\ref{D:#1}}
\newcommand{\refr}[1]{Remark ~\ref{R:#1}}
\newcommand{\refe}[1]{\eqref{E:#1}}
\newenvironment{thm}[1]%
{ \begin{Thm} \label{T:#1} \ifShowLabels \TeXref{T:#1} \fi }%
{ \end{Thm} }
\renewcommand{\th}[1]{\begin{thm}{#1} \sl }
\renewcommand{\eth}{\end{thm} }
\newenvironment{lemma}[1]%
{ \begin{Lem} \label{L:#1} \ifShowLabels \TeXref{L:#1} \fi }%
{ \end{Lem} }
\newcommand{\lem}[1]{\begin{lemma}{#1} \sl}
\newcommand{\end{lemma}}{\end{lemma}}
\newenvironment{propos}[1]%
{ \begin{Prop} \label{P:#1} \ifShowLabels \TeXref{P:#1} \fi }%
{ \end{Prop} }
\newcommand{\prop}[1]{\begin{propos}{#1}\sl }
\newcommand{\end{propos}}{\end{propos}}
\newenvironment{corol}[1]%
{ \begin{Cor} \label{C:#1} \ifShowLabels \TeXref{C:#1} \fi }%
{ \end{Cor} }
\newcommand{\cor}[1]{\begin{corol}{#1} \sl }
\newcommand{\end{corol}}{\end{corol}}
\newenvironment{conjec}[1]%
{ \begin{Conjec} \label{Conj:#1} \ifShowLabels \TeXref{C:#1} \fi }%
{ \end{Conjec} }
\newcommand{\conj}[1]{\begin{conjec}{#1} \sl }
\newcommand{\end{conjec}}{\end{conjec}}
\newenvironment{defeni}[1]%
{ \begin{Def} \label{D:#1} \ifShowLabels \TeXref{D:#1} \fi }%
{ \end{Def} }
\newcommand{\defe}[1]{\begin{defeni}{#1} \sl }
\newcommand{\end{defeni}}{\end{defeni}}
\newenvironment{remark}[1]%
{ \begin{Rem} \label{R:#1} \ifShowLabels \TeXref{#1} \fi }%
{ \end{Rem} }
\newcommand{\rem}[1]{\begin{remark}{#1}}
\newcommand{\end{remark}}{\end{remark}}
\newcommand{\Label}[1]%
{\label{#1}%
\ifShowLabels \TeXref{#1} \fi }%
\newcommand{\eq}[1]%
{\newline \ifShowLabels \TeXref{{E:#1}} \fi
\begin{equation} \label{E:#1}}
\newcommand{\end{equation}}{\end{equation}}
\newcommand{ \begin{proof} }{ \begin{proof} }
\newcommand{ \end{proof} }{ \end{proof} }
\newcommand{K\"ahler }{K\"ahler }
\newcommand\tiltet{\widetilde{\tet}}
\newcommand\F{\calF}
\newcommand\fo{\F{|_Z}\otimes o(Z)}
\begin{document}
\title{Cohomology of a Hamiltonian $T$-space with involution}
\author{Semyon Alesker}
\address{Equipe d'Analyse\\
Case 186\\
Universite Paris 6\\
place Jussieu\\
75252 Paris, Cedex 05\\
France}
\email{<EMAIL>}
\author{Maxim Braverman}
\address{Department of Mathematics\\
The Ohio State University \\
Columbus, OH 43210 \\
USA}
\email{<EMAIL>}
\begin{abstract}
Let $M$ be a compact symplectic manifold on which a compact torus $T$
acts Hamiltonialy with a moment map $\mu$. Suppose there exists a symplectic
involution $\tet:M\to M$, such that $\mu\circ\tet=-\mu$. Assuming that
0 is a regular value of $\mu$, we calculate the character of the
action of $\tet$ on the cohomology of $M$ in terms of the character of
the action of $\tet$ on the symplectic reduction $\mu^{-1}(0)/T$ of
$M$. This result generalizes a theorem of R.~Stanley, who considered
the case when $M$ was a toric variety and $\dim T=\frac12\dim_\RR{}M$.
\end{abstract}
\maketitle
\sec{introd}{Introduction}
In \cite{Stanley87}, R.~Stanley proved a lower bound for the number of faces of
a centrally symmetric simple polytope. The main ingredient of his proof is the
following result: Suppose $M$ is a toric variety corresponding to a simple
centrally symmetric polytope $P$ of dimension $n$. Then $M$ is acted upon by an
$n$-dimensional torus $T$. The symmetry on $P$ induces an involution $\tet:M\to
M$, such that
\eq{compat0}
\tet\circ t \ = \ t^{-1}\circ \tet, \qquad \text{for any} \quad
t\in T.
\end{equation}
Then $\tet$ acts on the cohomology $H^{2i}(M,\CC)$ of $M$. The result of
Stanley states that the character of this action is equal to
$\binom{n}{i}$.
In this paper, we extend the result of Stanley to an arbitrary compact
symplectic manifold $M$, which possesses a Hamiltonian action of a torus $T$ of
arbitrary dimension and an involution $\tet$, which preserves the symplectic
form and satisfies \refe{compat0}. In this situation, the moment map for the
action of $T$ can be chosen so that $\mu\circ\tet=-\tet$. Assuming that $0$ is a
regular value of $\mu$, we compute (cf. \reft{main}) the character of the action
of $\tet$ on the cohomology $H^*(M,\CC)$ in terms of the action of $\tet$ on the
symplectic reduction $M_0=\mu^{-1}(0)/T$ of $M$. Note that, if
$\dim{}T=\frac12\dim{}M$, so that $M$ is a toric variety, then $M_0$ is a point
and our formula reduces to the theorem of Stanley. Hence, we obtain, in
particular, a new, more geometric proof, of Stanley's result.
The proof of our main theorem is based on a study of the action of
$\tet$ on the equivariant cohomology of $M$. More precisely, we compute
the {\em graded character} $\chi_\tet(t)$ of this action (cf.
\refe{chi}) in two different ways.
First, we uses the well known formula $H^*_T(M)=H^*(M)\otimes H_T(pt)$
to express $\chi_\tet(t)$ in terms of the character of the action of
$\tet$ on $H^*(M)$, cf. \refp{HT=HxA}.
Our second computation (cf. \refp{chitet}) expresses $\chi_\tet(t)$ in terms of
the character of the action of $\tet$ on the cohomology of the symplectic
reduction $H^*(M_0)$. This is done in the following way: We consider the square
$f=\<\mu,\mu\>$ of the moment map as a Morse-Bott function on $M$. It is
equivariant with respect to the action of the semi-direct product
$\ZZ_2\ltimes{}T=\tilT$ on $M$ (here the action of $\ZZ_2$ is generated by
$\tet$). In \refl{tetBT} we show that the $\tilT$-equivariant Morse inequalities
for $f$ {\em with local coefficients} are, in fact, equalities and, hence,
may be used to calculate the equivariant cohomology of $M$. This generalizes a
result of Kirwan, \cite[Th.~5.4]{Kirwan84}. Comparing the above Morse {\em
equalities} with different local coefficients, we calculate $\chi_\tet(t)$ via
the character of the action of $\tet$ on $M_0$.
Comparison of the above two expressions for $\chi_\tet(t)$ leads to a
proof of our main theorem.
\subsection*{Contents}
The paper is organized as follows:
In \refs{main} we formulate and prove our
main result -- \reft{main}. The proof is based on two statements
(Propositions~\ref{P:HT=HxA} and \ref{P:chitet}), which are proven in
the later sections.
In \refs{Ex} we present some examples and applications of
\reft{main}. In particular, we show that it implies the result of
Stanley \cite{Stanley87}. We also discuss applications of our theorem
to flag varieties.
In \refs{HT=HxA} we study the action of the involution $\tet$
on the equivariant cohomology of $M$ and prove \refp{HT=HxA} (the
expression of $\chi_\tet(t)$ in terms of the action of $\tet$ on $H^*(M)$).
Finally, in \refs{Morse}, we use the equivariant Morse inequalities to
prove \refp{chitet} (the expression of $\chi_\tet(t)$ in terms of the
action of $\tet$ on $H^*(M_0)$).
\sec{main}{Main theorem}
\ssec{setting}{}
Let $(M,\ome)$ be a compact $2n$-dimensional symplectic manifold endowed with a
Hamiltonian action of a compact $k$-dimensional torus $T$. Let
$\grt^*\simeq\RR^k$ denote the dual space to the Lie algebra of $T$
and let $\mu:M\to \grt^*$ be the moment map for the action of $T$ on
$M$.
Let $\tet:M\to M$ be an involution of $M$ such that $\tet^*\ome=\ome$
and
\eq{compat}
\tet (t\cdot m) \ = \ (t^{-1})\cdot m,
\quad \text{for any} \quad t\in T, \ m\in M.
\end{equation}
We can and we will normalize the moment map (which is defined up to an additive
constant) so that
\eq{normal}
\mu\circ\tet=-\mu.
\end{equation}
In this situation $\tet$ acts naturally on the cohomology $H^i(M)=H^i(M;\CC)$ of
$M$ with complex coefficients. Let $H^i(M)^+$ (resp. $H^i(M)^-$) denote the
subspace of $H^i(M)$ fixed by $\tet$ (resp. the subspace of $H^i(M)$ on which
$\tet$ acts as a multiplication by $-1$). Set
\[
h^{i,\pm}=\dim_\CC{}H^i(M)^\pm.
\]
Suppose now that zero is a regular value for the moment map $\mu$. Then
$\mu^{-1}(0)$ is a smooth manifold which is invariant under the actions of $T$
and $\tet$. Moreover, the action of $T$ on $\mu^{-1}(0)$ is {\em locally free},
i.e. each point of $\mu^{-1}(0)$ has at most finite stabilizer in $T$. Hence,
the symplectic reduction $M_0:=\mu^{-1}(0)/T$ is an {\em orbifold}. Let
$H^i(M_0)$ denote the cohomology of $M_0$ with complex coefficients.
The involution $\tet$ preserves $\mu^{-1}(0)$ and, hence, acts on $M_0$ and
$H^i(M_0)$. Let $h^{i,+}_0$ (resp. $h^{i,-}_0$) denote the dimension of the
subspace of $\tet$ invariant vectors in $H^i(M_0)$ (resp. the dimension of the
subspace of the vectors on which $\tet$ acts by multiplication by $-1$).
Our principal result is the following
\th{main}
In the situation described above
\eq{main}
h^{i,+} \ - \ h^{i,-} \ = \
\sum_{j=0}^{\min\, (k,[i/2])}
\binom{k}{j}\, (h^{i-2j,+}_0 - h^{i-2j,-}_0),
\end{equation}
for any $i=0,1\nek 2n$.
\eth
\rem{main}
a. \ The conditions of the theorem imply that $\tet$ acts freely on
the set $C$ of fixed points of any subtorus $T'$ of $T$. Moreover,
$\tet$ acts freely on the set of connected components of $C$. It
follows from the fact that,
if $\mu'$ is the moment map for the $T'$-action, then
$\mu'\circ\tet=-\tet\circ\mu'$ and the restriction of $\mu'$ on any
connected component is a non-zero constant.
b. \ The theorem remains true if $M$ is a symplectic orbifold, rather than a
smooth manifold. The proof is just a bit more complicated than the one we
present is this paper.
c. \ Moreover, the theorem may be generalized to the case when $M$ has more
serious singularities (say, to the case when $M$ is a singular algebraic
manifold). In this case, the usual cohomology must be replaced by the
intersection cohomology. The use of the intersection cohomology also allows to
release the assumption that 0 is a regular value of the moment map. The details
will appear elsewhere.
d. \ The special case of the theorem, when $M$ is a toric manifold, is due to
R.~Stanley \cite{Stanley87}. In this sense, our result is a generalization of
Stanley's theorem. In particular, we obtain a new, more geometric proof, of the
Stanley's theorem. See \refss{Stanley} for details.
\end{remark}
\ssec{sketch}{}
The proof of the theorem is based on a study of the action of $\tet$ on the
equivariant cohomology of $M$. We now formulate the main results about this
action. The proofs are given on Sections~\ref{S:HT=HxA} and \ref{S:Morse}. Some
examples and applications of \reft{main} are discussed in \refs{Ex}.
\ssec{eqcoh}{Action of the involution on the equivariant cohomology}
Let $ET$ denote the universal $T$-bundle and let $BT=ET/T$ denote the
classifying space of $T$. Let $H_T^*(M)=H^*(ET\times_TM;\CC)$ denote the
equivariant cohomology of $M$ with complex coefficients. Then $\tet$ acts
naturally on $H_T^*(M)$, cf. \refss{action}. Let
\eq{chi}
\chi_\tet(t) \ := \ \sum_{t=0}^\infty\, t^i\Tr \tet|_{H_T^i(M)}
\end{equation}
denote the {\em graded character} of this action.
In \refs{HT=HxA} we use the fact that the spectral sequence of the fibration
$M\otimes_TET\to BT$ degenerates at the second term (cf. \cite[Proof of
Pr.~5.8]{Kirwan84}, \cite[Proof of Th.~5.3]{BrFar3}) to prove the following
proposition, in which we do not assume that 0 is a regular value of the moment
map.
\prop{HT=HxA}
\(
\displaystyle
\chi_\tet(t) \ = \
\sum_{i=1}^{2n} (h^{i,+} - h^{i,-}) \frac{t^{i}}{(1+t^2)^k}.
\)
\end{propos}
On the other side, in \refs{Morse} we use a version of the equivariant Morse
inequalities \cite{AtBott82} constructed in \cite{BrFar4,BrFar3} to get the
following
\prop{chitet}
Suppose that zero is a regular value for the moment map $\mu$ and let
$M_0:=\mu^{-1}(0)/T$. Then $\chi_\tet$ is equal to the graded character of the
action of $\tet$ on $H^*(M_0)$:
\eq{chitet}
\chi_\tet(t) \ = \ \sum_{i}\, (h^{i,+}_0 - h^{i,-}_0) t^i.
\end{equation}
\end{propos}
\cor{toric}
If, in the conditions of \refp{chitet}, the dimension of the torus is equal to
$n=\frac12\dim_\RR{M}$, then $\chi_\tet=1$.
\end{corol}
\ssec{prmain}{Proof of the main theorem}
Comparing Propositions~\ref{P:HT=HxA} and \ref{P:chitet} we obtain
\[
\sum_{i=1}^{2n} (h^{i,+} - h^{i,-}) \frac{t^{i}}{(1+t^2)^k} \ = \
\sum_{i}\, (h^{i,+}_0 - h^{i,-}_0) t^i,
\]
which is, obviously, equivalent to \reft{main}.
\hfill$\square$
\sec{Ex}{Examples and applications}
\ssec{Stanley}{Symmetric toric variety. Application to combinatorics}
\reft{main} takes a
particularly simple form if the dimension of the torus is equal to
$n=\frac12\dim_\RR M$, so that $M$ is a toric variety. In this case we
will say that $M$ is a {\em symmetric (with respect to the involution
$\tet$) toric variety}.
If $M$ is a symmetric toric variety, then
the reduced space $M_0$ is a point. Hence, $h^{i,-}_0=0$ for all $i$,
$h^{i,+}_0=0$ for all $i>0$ and $h^{0,+}_0=1$. The \reft{main} reduces
in this case to the following statement, which was originally proven
by R.~Stanley by a completely different method
\footnote{Stanley proved the theorem for more general case, when $M$
is a symmetric toric orbifold. One can prove this result using our
method and \refr{main}.b}.
\cor{Stanley} If, in the conditions of \reft{main}, the dimension of the
torus is equal to $n=\frac12\dim_\RR M$, then
\[
h^{i,+} \ - \ h^{i,-} \ = \ \binom{n}{i}.
\]
\end{corol}
There are a lot of examples of symmetric toric varieties. To describe these
examples let us recall that each toric variety is completely characterized by
its {\em moment polytope} $\mu(M)\subset \grt^*$. The toric variety $M$ is an
orbifold if and only if the polytope $\mu(M)$ is {\em simple}, i.e. if each of
its vertices has the valence $n=\frac12\dim_\RR M$. There is also a complete
description of polytopes corresponding to smooth toric varieties, cf., for
example, \cite[\S{}IV.2]{Audin91}. Now, the equation \refe{normal} implies that
the moment polytope corresponding to a symmetric toric variety is centrally
symmetric. Vice versa, if the moment polytope is centrally symmetric, one easily
constructs an involution on $M$ satisfying \refe{normal}. Hence, symmetric toric
orbifolds are in one-to-one correspondence with centrally symmetric simple
convex polytopes. \refc{Stanley} can be used to get an estimate on the number of
faces of such a polytope. See \cite{Stanley87} for details.
\rem{Campo} Recently, A. A\'{}Campo-Neuen \cite{Campo} extended
\refc{Stanley} to singular toric varieties. This leads to an extension of
Stanley's estimates on the number of faces of a centrally symmetric polytopes to
rational polytopes, which are not necessarily simple. This result may be also
obtained by our method, cf. \refr{main}.c.
\end{remark}
\ssec{CPn}{$\CC{P}^3$ as an $S^1$-space with involution}
Some toric varieties, which are not symmetric, still posses an
involution compatible, in the sense of \refe{compat}, with the action
of a torus of smaller dimension. Before discussing more general
examples, let us consider the simplest case $M=\CC{P}^{3}$. Then the
moment polytope is a $3$-dimensional simplex, which is, obviously, not
centrally symmetric. Hence, there is no involution on $M$ compatible,
in the sense of \refe{compat}, with the action of a $3$-dimensional
torus. However, using the homogeneous coordinates $[z_1:z_2:z_3:z_4]$
on $M$, we can define the action of the circle $T=S^1$ on $M$ and an
involution $\tet:M\to M$ by
\begin{gather}
t \ \cdot \ [z_1:z_2:z_3:z_4] \ = \ [tz_1:tz_2:z_3:z_4], \qquad
t\in S^1=\{\zet\in\CC:\, |\zet|=1\} \notag \\
\tet \ \cdot \ [z_1:z_2:z_3:z_4] \ = \ [z_3:z_4:z_1:z_2].\notag
\end{gather}
Clearly, all the conditions of \reft{main} are satisfied.
In this case, both sides of \refe{main} may be easily calculated. In particular,
$M_0=\CC{P}^1\times\CC{P}^1$ and $\tet$ acts on $M_0$ by
$\tet:(a,b)\mapsto{}(b,a)$, where $a,b\in\CC{P}^1$. We leave to the interested
reader to verify \reft{main} in this simple case.
\ssec{sym}{Toric varieties with involution}
The previous example allows the following generalization. Let $M$ be a
toric variety endowed with an action of the torus $T$ of dimension
$n=\frac12\dim_\RR M$. Suppose that there exists a linear involution
$\tiltet:\grt^*\to\grt^*$ which preserves the moment polytope
$\mu(M)$. Then $\tiltet$ induces an involution $\tet:M\to M$.
However, in general, this involution is not compatible with the
action of $T$ in the sense of \refe{compat}.
Let $\tiltet^*:\grt\to\grt$ denote the involution of the Lie algebra $\grt$ dual
to $\tiltet$. Let $T'\subset T$ be a subtorus, such that $\tiltet^*$ acts as
multiplication by $-1$ on the Lie algebra of $T'$. Then the actions of $T'$ and
$\tet$ on $M$ are compatible in the sense of \refe{compat}. If, in addition, 0
is a regular value of the moment map for the $T'$ action, then \reft{main} may
be applied.
\ssec{flag}{The variety of complete flags in $\CC^n$}
We now give an example of a manifold, which is not toric, but
satisfies the conditions of \reft{main}.
Let $\lam=\{\lam_1\nek\lam_n\}$ be a centrally symmetric set of real numbers. In
other words, $\lam=\{\pm\nu_1\nek\pm\nu_k\}$ if $n=2k$ and
$\lam=\{0,\pm\nu_1\nek\pm\nu_k\}$ if $n=2k+1$. We will assume that the numbers
$\nu_i$ above are positive and mutually different. The set $A_\lam$ of all complex
Hermitian $n\times{}n$-matrices with spectrum $\lam$ is naturally identified
with the variety of complete flags in $\CC^n$. In particular, $A_\lam$ has a
structure of a compact K\"ahler manifold.
Fix mutually different rational numbers $r_1\nek r_n$ and consider the action of
the circle $S^1=\{e^{it}:\, t\in\RR\}$ on $A_\lam$, defined by
\eq{actionF}
e^{it}\cdot A \ = \
\diag\big\{\, e^{ir_1t}\nek e^{ir_nt}\, \big\} \ \cdot \ A \
\cdot \
\diag\big\{\, e^{-ir_1t}\nek e^{-ir_nt}\, \big\},
\qquad t\in\RR, \ A\in A_\lam.
\end{equation}
This action is Hamiltonian with respect to the K\"ahler
structure on $A_\lam$: if we identify the coalgebra Lie of $S^1$ with
$\RR$, then the function
\[
\mu:\, A_\lam \ \to \RR, \qquad A=\{a_{ij}\} \ \mapsto \
\sum_{i=1}^n r_ia_{ii}
\]
is a moment map for this action.
Define an involution $\tet:A_\lam\to A_\lam$ by the formula
\[
\tet:\, A \ \mapsto \ -A^t.
\]
This involution preserves the K\"ahler structure on $A_\lam$ and satisfies
\refe{compat}, \refe{normal}. Moreover, one easily checks that $S^1$ acts
locally freely on $\mu^{-1}(0)$. Hence, zero is a regular value of $\mu$ and
\reft{main} is applicable.
\rem{flag}
a. \ The action \refe{actionF} is a restriction of the conjugate action on
$A_\lam$ of the torus of all unitary diagonal matrices. The latter action is
also Hamiltonian and compatible with the involution $\tet$ in the sense of
\refe{compat}. Unfortunately, zero is not a regular value of the moment map for
this action. However, the generalization of \reft{main}, indicated in
\refr{main}.c, may be applied in this case.
b. \ One easily generalizes the results of this subsection to a flag variety of
an arbitrary reductive group $G$. The involution $\tet$ should be replaced by
the action of the longest element of the Weyl group. Also the action of the
maximal torus of $G$ should be restricted to a subtorus on whose Lie algebra the
longest element of the Weyl group acts as a multiplication by $-1$, cf.
\refss{sym}. We leave the details to the interested reader.
\end{remark}
\sec{HT=HxA}{Proof of \refp{HT=HxA}}
\ssec{action}{Action of $\tet$ on $H_T^*(M)$}
Before proving the proposition let us describe more explicitly the
action of $\tet$ on $H_T^*(M)$. One of the ways to
define this action is the following. Consider the group
\[
\tilT \ := \ \big\{\, (\eps,t):\, \eps=\pm1, t\in T\, \big\}
\]
with the product law
\[
(\eps_1,t_1)\cdot (\eps_2,t_2) \ = \ (\eps_1\eps_2, t_1^{\eps_2}t_2).
\]
Then $\tilT$ acts on $M$ by the formula
\[
(\eps,t)\cdot m \ = \ t^\eps\cdot \tet^\eps(m).
\]
We will identify $\tet$ with the element $(-1,1)\in \tilT$. Let
$E\tilT$ denote the universal $\tilT$-bundle. Then $\tet$ acts
diagonally on $E\tilT\times_TM$ and, hence, on
$H_T^*(M)=H^*(E\tilT\times_TM)$ (in the last equality we used that
$E\tilT$ may be considered also as a model for the universal
$T$-bundle).
\ssec{tetBT}{Action of the involution on $BT$}
We consider the quotient $BT={E\tilT}/T$ as a model for the classifying space of
$T$. Then $\tet=(-1,1)\in\tilT$ acts on $BT$ and, hence, on the cohomology ring
$H^*(BT)$. The later is isomorphic to the graded ring $\CC[\grt]$ of polynomials
on the Lie algebra $\grt$ of $T$ (with grading given by twice the degree of the
polynomial):
\eq{isom}
H^*(BT) \ \simeq\CC[\grt].
\end{equation}
The following lemma describes the action of $\tet$ on this ring.
\lem{tetBT}
Under the isomorphism \refe{isom} the action of $\tet$ on $H^*(BT)$ is
given by the rule
\[
\tet(t_i) \ = \ -t_i.
\]
In other words, the restriction of $\tet$ on $H^2(BT)$ is equal to $-1$.
\end{lemma}
\begin{proof}
This lemma is well known, but in view of the difficulty we have to
locate an explicit reference, we present a proof here.
For simplicity, we only present the proof for the case $\dim_\RR{}T=1$.
The arguments in general case are exactly the same, but the notation
is more complicated. We also identify the one dimensional torus with
the unit circle $S^1=\{e^{i\phi}:\, \phi\in\RR\}$ in $\CC$.
Fix $m\ge3$ and consider the free action of $\tilT$ on the product of spheres
\[
S^{2m-1}\times S^{2m-1} \ := \ \big\{\,
(z_1\nek z_m;w_1\nek w_m)\in \CC^{2m}:\,
|z_1|^2+\cdots|z_m|^2=1;\, |w_1|^2+\cdots|w_m|^2=1
\big\},
\]
given by
\begin{gather}
e^{i\phi}\cdot(z_1\nek z_m;w_1\nek w_m) \ \mapsto
(e^{i\phi}z_1\nek e^{i\phi}z_m;w_1\nek w_m); \notag
\\
(-1,0)\cdot(z_1\nek z_m;w_1\nek w_m) \ \mapsto
(\oz_1\nek \oz_m;-w_1\nek -w_m). \notag
\end{gather}
Let us denote by $BT_m$ the quotient of $S^{2m-1}\times S^{2m-1}$ by the action
of $T\subset\tilT$. Since the reduced cohomology of $S^{2m-1}\times S^{2m-1}$
vanishes up to dimension $2m-2\ge4$ it follows that there is a natural $\tilT/T$
equivariant isomorphism between the cohomology $H^i(BT_m)$ and $H^i(BT)$ for any
$i$ less than $2m-3\ge 3$ (this may be shown by the same arguments as in the
proof of Lemma~2.8 in \cite{BrFar3})
Clearly, $\tet$ acts on $BT_m=\CC\PP^{m-1}\times S^{2m-1}$ as complex
conjugation on the first factor and multiplication by $-1$ on the second factor.
Hence, the induced action on the cohomology $H^2(BT_m)$ is multiplication by
$-1$. This proves the lemma.
\end{proof}
\cor{charBT}
The graded character of the action of $\tet$ on $H^*(BT)$ is equal to
$(1+t^2)^k$.
\end{corol}
\ssec{spseq}{The spectral sequence. Proof of \refp{chitet}}
The projection $E\tilT\times M\to E\tilT$ induces a $\tilT/T$
equivariant fiber bundle $p:E\tilT\times_TM\to BT$. The fiber of this
bundle is $\tilT/T$ equivariantly homeomorphic to $M$. The cohomology
$H^*(E\tilT\times_TM)=H^*_T(M)$ may be calculated by the spectral
sequence of the above bundle. The later spectral sequence degenerates
at the second term (cf. \cite[Proof of
Pr.~5.8]{Kirwan84}, \cite[Proof of Th.~5.3]{BrFar3}) and is obviously
$\tilT/T$ invariant. It follows that there exists a $\tilT/T$
equivariant isomorphism of graded rings
\[
H^*_T(M) \ \simeq \ H^*(M) \otimes_\CC H^*(BT).
\]
Hence, the graded character $\chi_\tet(t)$ of the action of $\tet$ on $H^*_T(M)$
is equal to the product of the graded characters of the actions of $\tet$ on
$H^*(M)$ and $H^*(BT)$. \refp{chitet} follows now from \refc{charBT}.
\hfill$\square$
\sec{Morse}{Equivariant Morse inequalities and proof of \refp{chitet}}
Our proof of \refp{chitet} is an application of the
$\tilT$-equivariant Morse inequalities for the square of the moment
map. Since the group $\tilT$ is disconnected, it is important to use
the equivariant Morse inequalities with local coefficients, cf.
\cite{BrFar4,BrFar3}.
The key fact of the proof is that the square of the moment map is an
{\em equivariantly perfect Morse function}, i.e. the above inequalities
are, in fact, equalities, cf. \refl{perfect}. This result is a slight
generalization of a theorem of Kirwan \cite[6, Th.~5.4]{Kirwan84}, who
considered $T$-equivariant Morse inequalities with trivial local
coefficients.
\ssec{morse}{Equivariant Morse inequalities}
For convenience of the reader we recall here equivariant Morse inequalities with
local coefficients for an action of a disconnected group, as they formulated in
\cite{BrFar4,BrFar3}. In this subsection $M$ is a compact manifold acted upon by
a compact, not necessarily connected, Lie group $G$. Let $\calF$ be a
$G$-equivariant flat vector bundle over $M$. Denote by $H^*_G(M,\calF)$ the
$G$-equivariant cohomology of $M$ with coefficients in $\calF$ and let
\[
\calP^G_\calF(t) \ = \ \sum_{i}\, t^i \dim_\CC\, H^i_G(M,\calF)
\]
be the {\em equivariant Poincar\'e series} of $M$ with coefficients on $\calF$.
Suppose $f:M\to\RR$ is a $G$-invariant function on $M$ and let $C$ denote the set of
critical points of $f$. We assume that $f$ is {\it non-degenerate in the sense of
Bott}, i.e. $C$ is a submanifold of $M$ and the Hessian of $f$ is a
non-degenerate on the normal bundle $\nu(C)$ to $C$ in $M$.
Let $Z$ be a connected component of the critical point set $C$ and let
$\nu(Z)$ denote the normal bundle to $Z$ in $M$. The bundle $\nu(Z)$
splits into the Whitney sum of two subbundles $\nu(Z) = \nu^+(Z)
\oplus \nu^-(Z)$, such that the Hessian is strictly positive on
$\nu^+(Z)$ and strictly negative on $\nu^-(Z)$. The dimension of the
bundle $\nu^-(Z)$ is called the {\it index} of $Z$ (as a critical
submanifold of $f$) and is denoted by $\ind(Z)$. Let $o(Z)$ denote
the {\it orientation bundle of $\nu^-(Z)$, considered as a flat line
bundle}.
If the group $G$ is connected, then $Z$ is a $G$-invariant submanifold
of $M$. In general, we denote by $G_Z=\{g\in G| \ g\cdot Z\subset Z\}$ the
stabilizer of the component $Z$ in $G$. Let $|G:G_Z|$ denote the
index of $G_Z$ as a subgroup of $G$; it is always finite.
The compact Lie group $G_Z$ acts on the manifold $Z$ and the flat
vector bundles $\F{|_Z}$ and $o(Z)$ are $G_Z$-equivariant. Let
$H_{G_Z}^\ast(Z,\fo)$ denote the equivariant cohomology of the
flat $G_Z$-equivariant vector bundle $\fo$. Consider the {\it
equivariant Poincar\'e series}
$$
\calP^{G_Z}_{Z,\F}(t)\ =\
\sum_i t^i \dim_{\CC}H_{G_Z}^i(Z,\fo)
$$
and define, using it, the following {\it equivariant Morse counting
series}
$$
\calM_{f,\F}^G(t)\ = \ \sum_Z
t^{\ind(Z)} |G:G_Z|^{-1}\calP_{Z,\F}^{G_Z}(t)
$$
where the sum is taken over all connected components $Z$ of $C$.
The following version of the equivariant Morse inequalities is a
particular case of \cite[Th.~7]{BrFar4},\cite[Th.~1.7]{BrFar3}.
\th{BrFar}
Suppose that $G$ is a compact Lie group, acting on a closed manifold
$M$, and let $\F$ be an equivariant flat vector bundle over $M$.
Then for any non-degenerate (in the sense of Bott) $G$-equivariant
function
$f:M\to\RR$, there exists a formal power series $\calQ(t)$ with
non-negative integer coefficients, such that
\[
\calM_{f,\F}^G(t) \ - \ \calP_{\F}^G(t) \ = \ (1+t)\calQ(t).
\]
\eth
\ssec{eqbundle}{Equivariant flat bundles}
We now return to the situation described in \refs{main}.
Let $\calF$ be a $\tilT$-equivariant flat vector bundle over $M$. This
is the same as $T$-equivariant flat vector bundle, on which $\tet$ acts
preserving the connection and so that $\tet(t\cdot \xi)=(t^{-1})\cdot\xi$ for
any $\xi\in\calF, t\in T$.
In our proof of \refp{chitet} we will only use the following type of
$\tilT$-equivariant bundles: Let $\rho:\ZZ_2\to \End_\CC(V_\rho)$ be
a representation of $\ZZ_2$ in a finite-dimensional complex vector
space $V_\rho$. Consider the bundle $\calF_\rho=M\times V_\rho$ with
the trivial connection and with the action of $\tilT$ given by
\[
(\eps,t):\, (m,\xi) \mapsto \big(\, (\eps,t)\cdot m,\, \rho(\eps)\cdot\xi\,
\big).
\]
Then the the equivariant cohomology $H_{\tilT}^*(M,\calF_\rho)$ of $M$ with
coefficients in $\calF_\rho$ is given by
\eq{HFrho}
H_{\tilT}^*(M,\calF_\rho) \ = \
\Hom_{\ZZ_2}\big(V_\rho^*, H_{T}^*(M)\big)
\end{equation}
(here $V^*_\rho$ is the representation dual to $V_\rho$). In particular, if
$\rho$ is the regular representation of $\ZZ_2$ then
\eq{reg}
H_{\tilT}^*(M,\calF_\rho)= H_{T}^*(M).
\end{equation}
The bundle $\calF_\rho$ is completely determined by the representation $\rho$.
When it causes no confusion, we will write $\rho$ for $\calF_\rho$ in oder to
simplify the notation.
\ssec{f}{Morse equalities for the square of the moment map}
Fix a scalar product $\<\cdot,\cdot\>$ on $\grt^*$ and consider the real-valued
function $f=\<\mu,\mu\>$ on $M$. Then $f$ is a $\tilT$ invariant Morse function
on $M$. Moreover, if $Z$ is any connected component of the set $C$ of critical
points of $f$, then the dimension of $Z$ is even and the bundles $\nu^{\pm}(Z)$
(cf. \refss{morse}) are orientable. In particularly, the orientation bundle
$o(Z)$ is trivial.
The set $Z_0:=\mu^{-1}(0)$ is a connected component of the set of critical
points of $f$. The group $\tilT$ acts on $Z_0$ and the equivariant
cohomology
\[
H^*_\tilT(Z_0,\calF_\rho|_{Z_0}) \ = \
\Hom_{\ZZ_2}(V^*_\rho,H^*(M_0)).
\]
The equivariant Poincar\'e series of $Z_0$ with coefficients in
$\calF_\rho$ is given by
\[
\calP^\tilT_{Z_0,\rho}(t) \ = \
\sum_i\, t^i\dim_\CC H^*_\tilT(Z_0,\calF_\rho|_{Z_0})
\ = \
\sum_i\, t^i\dim_\CC\Hom_{\ZZ_2}(V^*_\rho,H^*(M_0)).
\]
In particular,
\eq{Z0}\begin{aligned}
\calP^\tilT_{Z_0,\rho}(t) \ &= \ \sum_i\, t^ih^{i,+}_0,
\quad\text{if $\rho$ is the trivial representation}; \\
\calP^\tilT_{Z_0,\rho}(t) \ &= \ \sum_i\, t^ih^{i,-}_0,
\quad\text{if $\rho$ is the sign representation}.
\end{aligned}\end{equation}
The stabilizer $G_{Z_0}$ of $Z_0$ coincides with the whole group $\tilT$. Thus
$|G:G_{Z_0}|=1$.
The involution $\tet$ acts freely on the set of the connected components of $C$
different from $Z_0$, cf. \refr{main}.a. Hence, if $Z\not=Z_0$ is a connected
component of $C$, then $G_Z=T$ and $|G:G_Z|=2$. The $T$-equivariant Poincar\'e
polynomial of $Z$ is independent of $\rho$ and is given by
\[
\calP^T_{Z}(t) \ = \ \dim_\CC V_\rho\cdot
\sum_i\, t^i\dim_\CC H^i_T(Z,\calF|_Z).
\]
Hence, the equivariant Morse counting series
\[
\calM^{\tilT}_{f,\rho(t)} \ = \
\calP^\tilT_{Z_0,\rho}(t) \ + \ \frac12\sum_Z\, \calP^T_{Z}(t).
\]
Here the sum in the right hand side is taken over all connected components of
the set of critical points of $f$ different from $Z_0$.
Let
\[
\calP^\tilT_\rho(t) \ = \ \sum_i\, t^i \dim_\CC H^i_\tilT(M,\calF_\rho)
\]
be the equivariant Poincar\'e series of $M$ with coefficients in $\calF_\rho$.
The following lemma expresses the fact that $f$ is an {\em equivariantly perfect
Morse function}.
\lem{perfect}
The following equality holds
\eq{perfect}
\calM^{\tilT}_{f,\rho}(t) \ = \ \calP^\tilT_\rho(t).
\end{equation}
\end{lemma}
\begin{proof}
It follows from \reft{BrFar}, that there exists a formal power series
$Q_\rho(t)$ with non-negative coefficients, such that
\eq{morse}
\calM^{\tilT}_{f,\rho(t)} \ = \ \calP^\tilT_\rho(t) \ = \
\sum_i\, t^i H^i_\tilT(M,\calF_\rho) \ + \ (1+t)Q_\rho(t).
\end{equation}
Our goal is to show that $Q_\rho\equiv0$.
It follows from \refe{HFrho}, that both the equivariant Morse
counting series and the equivariant Poincar\'e series are additive with respect
to $\rho$. More precisely, if $\rho_1\oplus\rho_2$ denotes the direct sum of two
representations then
\[
\calM^{\tilT}_{f,\rho_1\oplus\rho_2}(t) \ = \
\calM^{\tilT}_{f,\rho_1}(t) \ + \ \calM^{f,\tilT}_{\rho_2}(t), \quad
\calP^\tilT_{\rho_1\oplus\rho_2}(t) \ = \
\calP^\tilT_{\rho_1}(t) \ + \ \calP^\tilT_{\rho_2}(t).
\]
Hence, it suffices to prove the lemma for the irreducible representations of
$\ZZ_2$. Moreover, it follows from \refe{morse} that it is enough to prove that
$Q_\rho=0$ when $\rho$ is a reducible representation, which contains any of the
irreducible representation as a subrepresentation.%
In particular, it is enough to prove that $Q_\rho=0$ when $\rho$ is
the regular representation. However, \refe{reg} implies that, if
$\rho$ is the regular representation, then \refe{morse} reduces to the
$T$-equivariant Morse inequalities with trivial coefficients. It was
shown by Kirwan \cite[Th.~5.4]{Kirwan84} that the later inequalities
are exact, i.e. $Q_\rho=0$. \end{proof}
\ssec{proof}{Proof of \refp{chitet}}
Set
\[
H^*_T(M)^\pm \ := \ \big\{\, x\in H^*_T(M):\, \tet x=\pm x\, \big\}
\]
Let us, first, apply \refl{perfect} with $\rho$ being the trivial representation
of $\ZZ_2$. It follows from \refe{HFrho} that, in this case,
$\calP^\tilT_{\rho}(t)=\sum_i\, t^i \dim_\CC H^i_T(M)^+$. Hence, from
\refe{Z0} and \refl{perfect} we obtain
\eq{1}
\sum_i\, t^i \dim_\CC H^i_T(M)^+ \ = \
\frac12\sum_Z\, \calP^T_{Z}(t) \ + \
\sum_i\, t^ih^{i,+}_0.
\end{equation}
Apply now \refl{perfect} with $\rho$ being the sign representation of $\ZZ_2$.
Then $\calP^\tilT_{\rho}(t)=\sum_i\, t^i \dim_\CC H^i_T(M)^-$ and
$\calP^\tilT_{Z_0,\rho}=\sum_i\, t^ih^{i,-}_0$. Hence,
\eq{2}
\sum_i\, t^i \dim_\CC H^i_T(M)^- \ = \
\frac12\sum_Z\, \calP^T_{Z}(t) \ + \
\sum_i\, t^ih^{i,-}_0.
\end{equation}
Subtracting \refe{2} from \refe{1} we get \refe{chitet}.
\hfill$\square$
\providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
|
\section{Introduction}
Deformation quantization was proposed about 22 years ago in the
pioneering work of Bayen, Flato, Fr\o nsdal, Lichnerowicz, and
Sternheimer [BFFLS] as an alternative to the usual correspondence
\begin{eqnarray*}
\mathrm{classical\ systems}&\leftrightarrow &\mathrm{quantum\
systems} \\
\mathrm{symplectic\ manifolds}&\leftrightarrow &\mathrm{Hilbert\ spaces}
\end{eqnarray*}
The idea is that algebras of observables in quantum mechanics
are ``close to" commutative algebras of functions on manifolds (phase spaces).
In other words, quantum algebras of observables are \textit{deformations}
of commutative algebras.
In the first order in perturbation theory one obtains automatically
a Poisson structure on phase space. Remind that a Poisson structure
on a smooth manifold $X$ is a bilinear operation $\{\cdot,\cdot\}$
on $C^{\infty}(X)$ satisfying the Jacobi identity
and the Leibniz rule with respect to the usual product in $C^{\infty}(X)$.
Two typical examples of Poisson manifolds are symplectic manifolds
and dual spaces to Lie algebras. A star-product on a Poisson manifold $X$
is an associative (but possibly non-commutative) product on $C^{\infty}(X)$
depending formally on a parameter usually denoted by~$\hbar$
(the ``Planck constant''). The product should have the form
\begin{equation}
f\star g= fg+\hbar\{f,g\}+\sum_{n\ge 2}\hbar^n B_n(f,g)
\end{equation}
where $(B_n)_{n\ge 1}$ are bidifferential operators
$C^{\infty}(X)\otimes C^{\infty}(X){\longrightarrow} C^{\infty}(X)$.
On the set of star-products acts an infinite-dimensional gauge group
of linear transformations of the vector space $C^{\infty}(X)$
depending formally on~$\hbar$, of the form
\begin{equation}
f\mapsto f +\sum_{n\ge 1} \hbar^n D_n(f)
\end{equation}
where the $(D_n)_{n\ge 1}$ are differential operators on $X$.
About two years ago I proved (see [K1]) that for every
Poisson manifold $X$ there is a canonically defined
gauge equivalence class of star-products on $X$.
This result gave a complete answer to first basic
questions in the program started by Mosh\'e Flato and co-authors.
I obtained the existence of a canonical deformation quantization
from a more general and stronger result, the formality theorem.
The statement of this theorem is that in a suitably defined
homotopy category of differential graded Lie algebras,
two objects are equivalent. The first object is the Hochschild complex
of the algebra of functions on the smooth manifold $X$, and the
second object is a $\mathbb{Z}$-graded Lie superalgebra of polyvector fields
on $X$. In the course of the proof I constructed an explicit isomorphism
in the homotopy category of Lie algebras for the case
$X=\mathbb{R}^n$. The terms in the formula for this isomorphism
can be naturally identified with Feynman diagrams
for certain two-dimensional quantum field theory
with broken rotational symmetry (see [CF] for a detailed
derivation of Feynman rules for this theory).
The moral is that a kind of ``string theory'' is necessary
for deformation quantization, which was originally
associated with quantum mechanics!
Mosh\'e was extremely happy when I told him about
my construction and we celebrated the ending of an old story
with a bottle of champagne in his Paris
apartment. Mosh\'e was very enthusiastic
about the new approach and expected new developments.
Later it became clear that not only there exists a
canonical way to quantize, but that one can define a universal
infinite-dimensional manifold parametrizing quantizations.
There are several evidences that this universal
manifold is a principal homogeneous space of the
so-called Grothen\-dieck-Teichm\"uller group, introduced by Drinfeld and
Ihara. At the ICM98 Congress in Berlin, I gave a talk
(which Mosh\'e qualified as ``wild") on relations between deformations,
motives and the Grothendieck-Teichm\"uller group. After the Congress
I decided not to write notes of the talk because one month later a
dramatic breakthrough in the area happened, which shed light on some parts
of my talk and required further work; the present paper fills this gap.
Dmitry Tamarkin found a new short derivation of the formality
theorem (for the case $X=\mathbb{R}^n$) from a very general result concerning
all associative algebras. Also from his result the (conjectured)
relation between deformation quantizations and the
Grothendieck-Teichm\"uller group seems to be much more
transparent. The present paper is a result of my attempts to
understand and generalize Tamarkin's results.
I use all the time the language of operads
and of homotopy theory of algebraic structures.
For me it seems to be the first real application of operads
to algebraic questions.
The main steps of the new proof of the formality theorem in
deformation quantization are the following:
\noindent STEP 1) On the cohomological Hochschild complex of any
associative algebra acts the operad of chains in the little discs operad
(Deligne's conjecture). This result is purely topological/combinatorial.
\noindent STEP 2)
The operad of chains of the little discs operad is formal, i.e.
it is quasi-isomorphic to its cohomology. More precisely, this is true
only in characteristic zero, e.g. over the field $\mathbb{Q}$ of rational numbers.
All the proofs of existence of a quasi-isomorphism
use certain multi-dimensional integrals and give explicit
formulas over real or over complex numbers.
\noindent STEP 3) From steps 1) and 2) follows that for any algebra $A$
over a field of characteristic zero, its cohomological
Hochschild complex $C^*(A,A)$ and its Hochschild cohomology $B:=H^*(A,A)$
are algebras over the \textit{same} operad (up to homotopy).
Moreover, if one chooses an explicit homotopy of complexes
between the Hochschild complex of $A$ and the graded space $B$ considered
as a complex with zero differential, one obtains \textit{two}
different structures of a homotopy Gerstenhaber algebra on $B$.
If these two structures are not equivalent, the first non-zero
obstruction to the equivalence gives a non-zero element in the
$H^1(\mathsf{Def}(B))$ of the deformation complex of the
Gerstenhaber algebra $B$.
\noindent STEP 4) For the case $A=\mathbb{R}[x_1,\dots,x_n]$, the algebra $B$
is the Gerstenhaber algebra of polynomial polyvector fields on $\mathbb{R}^n$.
An easy calculation shows that $H^1(\mathsf{Def}(B))$ coincides with
$B^2$ (i.e. with the space of bivector fields on $\mathbb{R}^n$).
The explicit homotopy between $B$ and $C^*(A,A)$ can be made invariant
under the group $\mathsf{Aff}(\mathbb{R}^n)$ of affine transformations of $\mathbb{R}^n$.
There is no non-zero $\mathsf{Aff}(\mathbb{R}^n)$-invariant bivector field on $\mathbb{R}^n$.
The conclusion is that two structures of the homotopy Gerstenhaber
algebra on $B$, mentioned in step 3), are equivalent.
In particular, these two structures give equivalent structures
of homotopy Lie algebra on $B$. This implies that
$C^*(A,A)$ is equivalent to $B$ as a homotopy Lie algebra,
which is the statement of the formality result in [K1].
Step 3) is the main discovery of D.~Tamarkin. It is an absolutely fundamental
new fact about \textit{all} associative algebras.
It implies in particular that the Hochschild complex is quasi-isomorphic
to another natural complex with strictly associative and commutative
cup-product. This other complex is hard to write down explicitly
at the moment. Also, there are many different
ways to identify operads as in step 2), even up to homotopy.
All these choices form an infinite-dimensional algebraic manifold defined
over the field of rational numbers $\mathbb{Q}$. On this manifold acts
the Grothendieck-Teichm\"uller group.
By the naturalness of the calculation in step 4), the same group
acts on quantizations of Poisson structures on $\mathbb{R}^n$, and also
on the set of gauge equivalence classes of star-products in $\mathbb{R}^n$.
Presumably, the action extends to general Poisson manifolds and to
general star-products.
The paper is organized as follows:
\noindent Section 2: an introduction to operads, and to
Deligne's conjecture concerning the Hochschild complex.
I would like to apologize for certain vagueness in subsections 2.5-2.7.
\noindent Section 3: a proof of the result of Tamarkin on the
formality of the chain operad of small discs operad, and a sketch of
its application to my formality theorem in deformation quantization.
In fact the idea of the proof presented here goes back to 1992-1993,
but somehow at that time I missed the point.
\noindent Section 4: I describe in elementary terms a version of
motives and of the motivic Galois group, and indicate its relations with
homogeneous spaces appearing in various questions in deformation
quantization.
\noindent Section 5: speculations about the possible r\^ole in Quantum
Field Theories of things described in previous sections.
\section{Deligne's conjecture and its generalization \\ to higher
dimensions}
\subsection{Operads and algebras}
Here I remind the definitions of an operad and of an algebra over an
operad (see also [GJ], [GK]). The language of operads is convenient
for descriptions and constructions of various algebraic structures.
It became quite popular in theoretical physics during the past few years
because of the emergence of many new types of algebras related with
quantum field theories.
\begin{definition}
An operad (of vector spaces) consists of the following:
\noindent 1) a collection of vector spaces $P(n)$, $n\ge 0$,
\noindent 2) an action of the symmetric group $S_n$ on $P(n)$ for every $n$,
\noindent 3) an identity element $\mathrm{id}_P\in P(1)$,
\noindent 4) compositions $m_{(n_1,\dots, n_k)}$:
\begin{equation}
P(k)\otimes \bigl(P(n_1)\otimes P(n_2)\otimes\dots\otimes
P(n_k)\bigr) {\longrightarrow} P(n_1+\dots+n_k)
\end{equation}
for every $k\ge 0$ and $n_1,\dots,n_k\ge 0$
satisfying a natural list of axioms which will be clear from examples.
\end{definition}
The simplest example of operad is given by
$P(n):= \mathsf{Hom}(V^{\otimes n}, V)$ where $V$ is a vector space.
The action of the symmetric group and the identity element are obvious,
and the compositions are defined by the substitutions
$$\bigl(m_{(n_1,\dots, n_k)}\bigl(\phi\otimes (\psi_1\otimes \psi_2\otimes
\dots \otimes \psi_k)\bigr)
\bigr)(v_1\otimes\dots\otimes v_{n_1+\dots +n_k})$$
$$:= \phi\bigl(\psi_1(v_1\otimes\dots \otimes v_{n_1})\otimes\dots
\otimes \psi_k(v_{n_1+\dots+n_{k-1}+1}\otimes\dots
\otimes v_{n_1+\dots + n_k})\bigr)$$
where $\phi\in P(k)=\mathsf{Hom}(V^{\otimes k},V),$ $\psi_i\in
P(n_i)=\mathsf{Hom}(V^{\otimes n_i},V)$, $i=1,\dots,k$.
This operad is called the endomorphism operad of a vector space. The
axioms in the definition of operads express natural properties of this
example. Namely, there should be an associativity law for multiple
compositions, various compatibilities for actions of symmetric groups, and
evident relations for compositions including the identity element.
Another important example of an operad is $\mathsf{Assoc}_1$.
The $n$-th component $\mathsf{Assoc}_1(n)$ for $n\ge 0$ is defined as
the collection of all universal ( = functorial) $n$-linear operations
$A^{\otimes n}{\longrightarrow} A$ defined on all associative algebras $A$ with unit.
The space $\mathsf{Assoc}_1(n)$ has dimension $n!$, and is spanned
by the operations
$$a_1\otimes a_2\otimes\dots\otimes a_n\mapsto a_{\sigma(1)}a_{\sigma(2)}
\dots a_{\sigma(n)}$$
where $\sigma\in S_n$ is a permutation. The space $\mathsf{Assoc}_1(n)$
can be identified with the subspace of the free associative unital algebra
in $n$ generators consisting of expressions polylinear in each generator.
\begin{definition}
An algebra over an operad $P$ consist of a vector space
$A$ and a collection of polylinear maps $f_n:P(n)\otimes A^{\otimes n}{\longrightarrow} A$
for all $n\ge 0$ satisfying the following list of axioms:
\noindent 1) for any $n\ge 0$ the map $f_n$ is $S_n$-equivariant,
\noindent 2) for any $a\in A$ we have $f_1(\mathrm{id}_P\otimes a)=a$,
\noindent 3) all compositions in $P$ map to compositions
of polylinear operations on $A$.
\end{definition}
In other words, the structure of algebra over $P$ on a vector space $A$
is given by a homomorphism of operads from $P$ to the endomorphism
operad of $A$. Another name for algebras over $P$ is $P$-algebras.
For example, an algebra over the operad $\mathsf{Assoc}_1$ is an associative
unital algebra. If we replace the $1$-dimensional space $\mathsf{Assoc}_1(0)$
by the zero space $0$, we obtain an operad $\mathsf{Assoc}$ describing
associative algebras possibly without unit. Analogously, there is an
operad denoted $\mathsf{Lie}$, such that $\mathsf{Lie}$-algebras are
Lie algebras. The dimension of the $n$-th component $\mathsf{Lie}(n)$ is
$(n-1)!$ for $n\ge 1$ and $0$ for $n=0$.
Let us warn the reader that not all algebraic structures correspond
to operads. Two examples of classes of algebraic structures that cannot
be cast in the language of operads are the class of fields and the
class of Hopf algebras.
We conclude this section with an explicit description of free algebras in
terms of operads.
\begin{theorem}
Let $P$ be an operad and $V$ be a vector space. Then the free
$P$-algebra $\mathsf{Free}_P(V)$ generated by $V$ is naturally isomorphic as a
vector space to
$$\bigoplus_{n\ge 0} \bigl(P(n)\otimes V^{\otimes n}\bigr)_{S_n}$$
\end{theorem}
The subscript $\cdot_{S_n}$ denotes the quotient space of coinvariants for
the diagonal action of group $S_n$. The free algebra $\mathsf{Free}_P(V)$
is defined by the usual categorical adjunction property:
{\it the set $\mathsf{Hom}_{\,P-\mathrm{algebras}}
(\mathsf{Free}_P(V),A)$ (i.e. homomorphisms in the category of $P$-algebras)
is naturally isomorphic to the set
$\mathsf{Hom}_{\,\mathrm{vector\,\,spaces}} (V,A)$ for any $P$-algebra $A$.}
\subsection{Topological operads, operads of complexes, etc.}
In the definition of operads we can replace vector spaces by
topological spaces, and the operation of tensor product by the usual
(Cartesian) product.
\begin{definition}
A topological operad consists of the following:
\noindent 1) a collection of topological spaces $P(n)$, $n\ge 0$,
\noindent 2) a continuous action of the symmetric group
$S_n$ on $P(n)$ for every $n$,
\noindent 3) an identity element $\mathrm{id}_P\in P(1)$,
\noindent 4) compositions $m_{(n_1,\dots, n_k)}$:
$P(k)\times \bigl(P(n_1)\times P(n_2)\times\dots\times
P(n_k)\bigr) {\longrightarrow} P(n_1+\dots+n_k)$ \\
which are continuous maps for every $k\ge 0$ and $n_1,\dots ,n_k\ge 0$
satisfying a list of axioms analogous to the one in the definition of an
operad of vector spaces.
\end{definition}
An analog of the endomorphism operad is the following one:
for any $n\ge 0$ the topological space $P(n)$ is the space of continuous
maps from $X^n$ to $X$, where $X$ is a given compact topological space.
In general, the definitions of an operad and of an algebra over an operad
can be made in arbitrary symmetric monoi\-dal category $\mathcal{C}$
(i.e in a category endowed with the functor $\otimes:\mathcal{C}\times \mathcal{C}{\longrightarrow}\mathcal{C}$,
the identity element $1_{\mathcal{C}}\in \mathsf{Objects}(\mathcal{C})$, and various coherence
isomorphisms for associativity, commutativity of $\otimes$, etc., see,
for example, [ML]).
We shall here consider mainly operads in the symmetric monoidal
category $\mathsf{Complexes}$ of $\mathbb{Z}$-graded complexes of abelian groups
(or vector spaces over a given field). Often operads in the category of
complexes are called \textit{differential graded operads}, or simply
\textit{dg-operads}. Each component $P(n)$ of an operad of complexes is a
complex, i.e. a vector space decomposed into a direct sum
$P(n)=\oplus_{i\in \mathbb{Z}} P(n)^i$, and endowed with a differential
$d:P(n)^i{\longrightarrow} P(n)^{i+1}$, $d^2=0$, of degree $+1$.
There is a natural way to construct an operad of complexes from a
topological operad by using a version of the singular chain complex.
Namely, for a topological space $X$, denote by $\mathsf{Chains}(X)$
the complex concentrated in nonpositive degrees, whose $(-k)$-th component
for $k=0,1,\dots$ consists of the formal finite additive combinations
$$\sum_{i=1}^N n_i \cdot f_i,\,\,\,n_i\in \mathbb{Z},\,\,\,N\in \mathbb{Z}_{\ge 0}$$
of continuous maps $f_i:[0,1]^k{\longrightarrow} X$ (``singular cubes'' in $X$),
modulo the following relations
\noindent 1) $f\circ \sigma=\mathit{sign}(\sigma) \, f$ for
any $\sigma \in S_k$ acting on the standard cube
$[0,1]^k$ by permutations of coordinates,
\noindent 2) $f'\circ pr_{k{\longrightarrow} (k-1)}=0$ where
$pr_{k{\longrightarrow} (k-1)}:[0,1]^k{\longrightarrow}[0,1]^{k-1}$ is the projection onto first
$(k-1)$ coordinates, and $f':[0,1]^{k-1}{\longrightarrow} X$ is a continuous map.
The boundary operator on cubical chains is defined in the usual way.
The main advantage of cubical chains with respect to simplicial
chains is that there is an external product map
\begin{equation}
\bigotimes_{i\in I} \bigl(\mathsf{Chains}(X_i)\bigr){\longrightarrow}
\mathsf{Chains}\bigl(\prod_{i\in I} X_i\bigr)
\end{equation}
which is a \textit{natural} homomorphism of complexes for any finite
collection $(X_i)_{i\in I} $ of topological spaces.
Now if $P$ is a topological operad then the collection of complexes
$\bigl(\mathsf{Chains}(P(n))\bigr)_{n\ge 0}$ has a natural operad
structure in the category of complexes of Abelian groups. The compositions
in $\mathsf{Chains}(P)$ are defined using the external tensor product
of cubical chains.
Passing from complexes to their cohomology we obtain an operad
$H_*(P)$ of $\mathbb{Z}$-graded Abelian groups ( = complexes with zero differential),
the homology operad of $P$.
\subsection{Operad of little discs}
Let $d\ge 1$ be an integer. Denote by $G_d$ the $(d+1)$-dimensional Lie
group acting on $\mathbb{R}^d$ by affine transformations $u\mapsto \lambda u +v$
where $\lambda>0$ is a real number and $v\in \mathbb{R}^d$ is a vector.
This group acts simply transitively on the space of closed discs in
$\mathbb{R}^d$ (in the usual Euclidean metric). The disc with center $v$ and
with radius $\lambda$ is obtained from the standard disc
$$D_0:=\{(x_1,\dots,x_d)\in \mathbb{R}^d | \,x_1^2+\dots +x_d^2\le 1\}$$
by a transformation from $G_d$ with parameters $(\lambda,v)$.
\begin{definition}
The little discs operad $C_d$ is a topological operad with the
following structure:
\noindent 1) $C_d(0)=\emptyset$,
\noindent 2) $C_d(1)= \,\mathrm{point}\,=\{\mathrm{id}_{C_d}\}$,
\noindent 3) for $n\ge 2$ the space $C_d(n)$ is the space of configurations
of $n$ disjoint discs $(D_i)_{1\le i\le n}$ inside the standard disc $D_0$.
The composition $C_d(k)\times C_d(n_1)\times\dots\times C_d(n_k){\longrightarrow}
C_d(n_1+\dots+n_k)$ is obtained by applying elements from
$G_d$ associated with discs $(D_i)_{1\le i\le k}$ in the configuration
in $C_d(k)$ to configurations in all $C_d(n_i),\,\,i=1,\dots,k$
and putting the resulting configurations together. The action of
the symmetric group $S_n$ on $C_d(n)$ is given by renumerations of
indices of discs $(D_i)_{1\le i\le n}$.
\end{definition}
The operad $C_d$ was introduced in 70-ies by Boardmann and Vogt,
and by Peter May (see [BV], [M]) in order to describe homotopy types of
$d$-fold loop spaces (i.e. spaces of continuous maps
$$\mathsf{Maps}((S^d,{\rm base \,\,\, point}), (X,x))$$
where $X$ is a topological space with base point $x$). It is the most
important operad in homotopy theory. Strictly speaking, topologists use a
slightly different model called the operad of little cubes, but the
difference is not essential because homotopically there is no difference
between cubes and discs.
The space $C_d(n)$ is homotopy equivalent to the configuration space
of $n$ pairwise distinct points in $\mathbb{R}^d$:
\begin{displaymath}
\mathsf{Conf}_n(\mathbb{R}^d):=\quad \left(\mathbb{R}^d\right)^n\setminus \mathrm{Diag}=
\{(v_1,\dots,v_n)\in \left(\mathbb{R}^d\right)^n |v_i\ne v_j
\ \mathrm{for\ any}\ i\ne j\}
\end{displaymath}
There is an obvious continuous map $C_d(n){\longrightarrow} \mathsf{Conf}_n(\mathbb{R}^d)$ which
associates a collection of disjoint discs with the collection of
their centers. This map induces a homotopy equivalence because each fiber
of this map is contractible.
The space $\mathsf{Conf}_2(\mathbb{R}^d)$ (and hence $C_d(2)$) is homotopy
equivalent to the $(d-1)$-dimensional sphere $S^{d-1}$.
The homotopy equivalence is given by the map
$$(v_1,v_2)\mapsto {v_1-v_2\over |v_1 -v_2|}\in S^{d-1}\subset \mathbb{R}^d$$
\subsection{Hochschild complex and Deligne's conjecture }
In 1993 Pierre Deligne made a conjecture relating the little discs operad
(in dimension $d=2$) and the cohomological Hochschild complex
of an arbitrary associative algebra $A$ (defined over any field $k$).
Namely, the Hochschild complex $C^*(A,A)$ is concentrated in non-negative
degrees and is defined as
\begin{equation}
C^n(A,A):=\mathsf{Hom}_{\mathrm{vector\,spaces}} (A^{\otimes n}, A),
\,\,\,n\ge 0
\end{equation}
and the differential in $C^*(A,A)$ is given by the formula
\begin{eqnarray*}
&&(d \phi)(a_1\otimes \dots \otimes a_{n+1}):=\\
&&a_1 \phi(a_2\otimes\dots
\otimes a_n)+\sum_{i=1}^n (-1)^i \phi(a_1\otimes
\dots\otimes a_i a_{i+1}\otimes \dots\otimes a_{n+1})+\cdots\\
&&\quad \cdots +(-1)^{n+1} \phi(a_1\otimes \dots\otimes a_n) a_{n+1}
\end{eqnarray*}
for any $\phi\in C^n(A,A)$.
The Hochschild complex plays a fundamental r\^ole in the deformation theory
of an associative algebra $A$. There are two basic operations on
$C^*:=C^*(A,A)$, the cup product $\cup:C^k\otimes C^l{\longrightarrow} C^{k+l}$
and the Gerstenhaber bracket $[\,,\,]:C^k\otimes C^l{\longrightarrow} C^{k+l-1}$.
The formulas for these operations are the following (here $\phi\in C^k$,
$\psi\in C^l$):
\begin{eqnarray*}
&&(\phi\cup\psi)(a_1\otimes\dots\otimes a_{k+l})\\
&&\quad :=(-1)^{kl}
\phi(a_1{\otimes\dots\otimes} a_k)\cdot\psi(a_{k+1}{\otimes\dots\otimes} a_{k+l})\\
&&[\phi,\psi]:=\phi\circ\psi- (-1)^{(k-1)(l-1)}
\psi\circ\phi\quad \mathrm{where}\\
&&(\phi\circ\psi)(a_1{\otimes\dots\otimes} a_{k+l-1}):=\\
&&\quad \sum_{i=1}^{k-1}(-1)^{i(l-1)}
\phi(a_1{\otimes\dots\otimes} a_i\otimes\psi(a_{i+1}{\otimes\dots\otimes} a_{i+l}) {\otimes\dots\otimes} a_{k+l-1})
\end{eqnarray*}
The Gerstenhaber bracket gives (after a shift of the $\mathbb{Z}$-grading
by~$1$) the structure of differential graded Lie algebra on the
Hochschild complex. The cup product has also a remarkable property:
it is \textit{not} graded commutative (it is associative only), but
the induced operation on the cohomology \textit{is} graded commutative.
Moreover, the Gerstenhaber bracket induces an operation on $H^*(A,A)$
which satisfies the Leibniz rule with respect to the cup product:
$$[\phi,\psi_1\cup \psi_2]=[\phi,\psi_1]\cup\psi_2 +(-1)^{(k-1)(l_1-1)}
\psi_1\cup [\phi,\psi_2],\,\,\,\phi\in C^k,\,\psi_i \in C^{l_i}$$
although this identity does not hold on the level of cochains.
The cohomology space $H^*(A,A)$ carries the structure
of the Gerstenhaber algebra, i.e. it is a graded vector space endowed
with a Lie bracket of degree $(-1)$ (satisfying the skew symmetry
condition and the Jacobi identity with appropriate signs),
and with a graded commutative associative product of degree $0$,
satisfying the graded Leibniz rule with respect to the bracket.
It was observed by several people (F.~Cohen in [C], P.~Deligne,...)
that the $\mathbb{Z}$-graded operad $\mathsf{Gerst}$ describing Gerstenhaber algebras
has a very beautiful topological meaning. Namely, it is naturally
equivalent to the homology operad of the topological operad $C_2$. The space
of binary operations $C_2(2)$ is homotopy equivalent to the circle $S^1$. The
Gerstenhaber bracket corresponds to the generator of $H_1(S^1)\simeq \mathbb{Z}$
and the cup product corresponds to the generator of $H_0(S^1)\simeq \mathbb{Z}$.
\begin{conjecture}[P.~Deligne]
There exists a natural action of the operad $\mathsf{Chains}(C_2)$ on the
Hochschild complex $C^*(A,A)$ for arbitrary associative algebra $A$.
\end{conjecture}
The story of this conjecture is quite dramatic. In 1994 E.~Getzler and
J.~Jones posted on the e-print server a preprint [GJ] in which the proof
of the Deligne conjecture was contained.
Essentially at the same time M.~Gerstenhaber and A.~Voronov in [GV] made
analogous claims. The result was considered as well established and was
actively used later. But in the spring of 1998 D.~Tamarkin observed that
there was a serious flaw in both preprints. The cell decomposition used
there turned out to be not compatible with the operad structure; the first
example of wrong behavior appears for operations with $6$ arguments.
I think that I have now a complete proof of the Deligne conjecture
(and its generalization, see the next section).
A combination of two results of Tamarkin (see [T1] and [T2])
also implies Deligne conjecture. Sasha Voronov says that he corrected
the problem in his approach, and there is also an announcement [MS] by
J.~McClure and J.~Smith with the same result.
Unfortunately, all ``proofs'' and announcements of proof are still too
complicated to be put here. We need a really short and convincing argument
for this very fundamental fact about Hochschild complexes.
It seems that the simplicity of the Hochschild complex is deceiving.
\subsection{Higher-dimensional generalization of Deligne's
conjecture}
I propose here an ``explanation'' of Deligne's conjecture and its natural
generalization.
The operad $\mathsf{Assoc}$ has a topological origin, it is the
homology operad of the little intervals (i.e. $1$-dimensional discs)
operad. Namely, for $n\ge 1$, the spaces $C_1(n)$ have $n!$ connected
components corresponding to permutations $\sigma\in S_n$, and each of
the components is contractible.
We can go still one step down, defining the operad $C_0$ as a really
trivial operad:
$$C_0(n)=\emptyset\,\,\,\,{\rm for}\,\,\,n\ne
1,\,\,\,\,\,C_0(1)=\mathrm{point}=\{\mathrm{id}_{C_0}\}\,\,$$
This definition is natural because one cannot put $\ge 2$ disjoint
zero-dimensional discs inside one zero-dimensional disc ( = point).
\begin{definition}
For $d\ge0$, a $d$-algebra is an algebra over the operad
$\mathsf{Chains}(C_d)$ in the category of complexes.
\end{definition}
For $d>0$ this notion was introduced by Getzler and Jones.
By definition, a $0$-algebra is just a complex. The notion of
$1$-algebra is very close to the notion of an associative algebra.
There is a well-known companion of associative algebras, a class of
so called $A_{\infty}$-algebras. One can show that any $1$-algebra carries
a natural structure of $A_{\infty}$-algebra, and from the points of view of
homotopy theory and of deformation theory, there is no difference
between associative algebras, $A_{\infty}$-algebras and $1$-algebras
(see 2.6 and 3.1). In particular, one can introduce the cohomological
Hochschild complex for $A_{\infty}$-algebras and $1$-algebras.
This Hochschild complex carries a natural structure of a differential
graded Lie algebra.
An $A_{\infty}$-version of Deligne's conjecture says that this Hochschild
complex carries naturally a structure of $2$-algebra, extending the
structure of differential graded Lie algebra. It has a baby version
in dimension $(0+1)$: if $A$ is a vector space (i.e. $0$-algebra
concentrated in degree 0) then the Lie algebra of the group of
affine transformations
$$\mathsf{Lie}(\mathsf{Aff}(A))=\mathsf{End}(A)\oplus A$$
has also a natural structure of an associative algebra, in particular it is
a $1$-algebra. The product in $\mathsf{End}(A)\oplus A$ is given by the
formula
$$(\phi_1,a_1)\times (\phi_2,a_2):=(\phi_1 \phi_2, \phi_1 (a_2))\,\,\,.$$
The space $\mathsf{End}(A)\oplus A$ plays the r\^ole of the Hochschild
complex in the case $d=0$ (see 2.7).
The definition of a $d$-algebra given above seems to be too complicated
for doing concrete calculations. At the end of section 3.2 we describe
much smaller operads which are quasi-isomorphic to the chain operads
of little discs operads. The reader will see that in dimension
$d\ge 2$ the situation is very simple.
Now we introduce the notion of \textit{action} of a $(d+1)$-algebra
on a $d$-algebra. It is convenient to formulate it using so-called
colored operads. Instead of defining exactly what a colored operad is,
we give one typical example: there is a colored operad with two colors such
that algebras over this operad are pairs $(\mathfrak{g},A)$ where $\mathfrak{g}$ is a Lie algebra
and $A$ is an associative algebra on which $\mathfrak{g}$ acts by derivations.
Let us fix a dimension $d\ge 0$. Denote by $\sigma:\mathbb{R}^{d+1}{\longrightarrow} \mathbb{R}^{d+1}$
the reflection
$$(x_1,\dots,x_{d+1})\mapsto (x_1,\dots,x_d,-x_{d+1})$$ at the coordinate
hyperplane, and by $\mathcal{H}_+$ the upper-half space
$$\{(x_1,\dots,x_{d+1})|\,x_{d+1}> 0\}$$
\begin{definition}
For any pair of non-negative integers $(n,m)$ we define a topological space
$SC_d(n,m)$ as
\noindent 1) the empty space $\emptyset$ if $n=m=0$,
\noindent 2) the one-point space if $n=0$ and $m=1$,
\noindent 3) in the case $n\ge 1$ or $m\ge 2$,
the space of configurations of $m+2n$ disjoint discs
$(D_1,\dots,D_{m+2n})$ inside the standard disc $D_0\subset \mathbb{R}^{d+1}$
such that $\sigma(D_i)=D_i$ for $i\le m$, $\sigma(D_i)=D_{i+n}$ for
$m+1\le i\le m+n$ and such that all discs $D_{m+1},\dots,D_{m+n}$ are
in the upper half space $\mathcal{H}_+$.
\end{definition}
The reader should think about points of $SC_d(n,m)$ as about configurations
of $m$ disjoint semidiscs $(D_1\cap \mathcal{H}_+,\dots,D_m\cap\mathcal{H}_+)$ and of
$n$ discs $(D_{m+1},\dots,D_{m+n})$ in the standard semidisc $D_0\cap\mathcal{H}_+$.
The letters ``$SC$'' stand for ``Swiss Cheese'' [V].
Notice that the spaces $SC_d(0,m)$ are naturally isomorphic to
$C_d(m)$ for all $m$. One can define composition maps analogously to
the case of the operad $C_d$:
$$SC_d(n,m)\times \bigl(C_{d+1}(k_1)\times\cdots\times C_{d+1}(k_n)\bigr)
\times\bigl(SC_d(a_1,b_1)\times\cdots\times SC_d(a_m,b_m)\bigr)$$
$${\longrightarrow} SC_d(k_1+\cdots+k_n+a_1+\cdots+a_m,b_1+\cdots+b_m)$$
\begin{definition}[A~.Voronov [V{]}]
The colored operad $SC_d$ has two colors and consists of collections
of spaces
$$\bigl(SC_d(n,m)\bigr)_{n,m\ge 0},\quad \bigl(C_{d+1}(n)\bigr)_{n\ge 0},$$
and appropriate actions of symmetric groups, identity elements, and
of all composition maps.
\end{definition}
As before, we can pass from a colored operad of topological spaces to
a colored operad of complexes using the functor $\mathsf{Chains}$.
\begin{definition}
An action of a $(d+1)$-algebra $B$ on a $d$-algebra $A$ is, on the pair
$(B,A)$, a structure of algebra over the colored operad
$\mathsf{Chains}(CS_d)$, compatible with the structures of algebras on
$A$ and on $B$.
\end{definition}
The \textit{generalized Deligne conjecture} says that for every
$d$-algebra $A$ there exists a universal (in an appropriate sense,
up to homotopy, see 2.6) $(d+1)$-algebra acting on $A$.
I think that I have a proof of this conjecture, so I am making the following
\begin{claim}
Let $A$ be a $d$-algebra for $d\ge 0$.
Consider the homotopy category of pairs $(B,\rho)$ where $B$ is a
$(d+1)$-algebra and $\rho$ is an action of $B$ on $A$.
In this category there exists a final object.
\end{claim}
Using this claim one can give the
\begin{definition}
For a $d$-algebra $A$ the (generalized) Hochschild complex
$\mathsf{Hoch}(A)$ is defined as the universal $(d+1)$-algebra acting on $A$.
\end{definition}
The universality of $\mathsf{Hoch}(A)$ means that in the sense
of homotopy theory of algebras, an action of $(d+1)$-algebra
$B$ on $A$ is the same as a homomorphism of $(d+1)$-algebras
$B{\longrightarrow} \mathsf{Hoch}(A)$.
In the next subsection we review homotopy and deformation theory of
algebras over operads, and in the following one we describe an
``explicit'' model for $\mathsf{Hoch}(A)$.
\subsection{Homotopy theory and deformation theory}
Let $P$ be an operad of complexes, and $f:A{\longrightarrow} B$ be a morphism of two
$P$-algebras.
\begin{definition}
$f$ is a quasi-isomorphism iff it induces an isomorphism of the cohomology
groups of $A$ and $B$ considered just as complexes.
\end{definition}
Two algebras $A$ and $B$ are called \textit{homotopy equivalent}
iff there exists a chain of quasi-isomorphisms
\begin{equation}
A=A_1{\longrightarrow} A_2{\longleftarrow} A_3{\longrightarrow} \dots{\longleftarrow} A_{2k+1}=B
\end{equation}
One can define a new structure of category on the collection of
$P$-algebras in which quasi-isomorphic algebras become equivalent.
There are several ways to do it, using either Quillen's machinery
of homotopical algebra (see [Q]), or using a free resolution of
the operad $P$, or some simplicial constructions, etc.
For example, in the category of differential graded Lie algebras,
morphisms in the homotopy category are so called $L_{\infty}$-morphisms
(see e.g. [K1]), modulo a suitably defined defined equivalence relation
(a homotopy between morphisms).
I shall not give here any precise definition of the homotopy
category in general, just say that morphisms in the homotopy category
of $P$-algebras are connected components of certain topological spaces,
exactly as in the usual framework of homotopy theory (i.e. in the
category of topological spaces).
In the case when the operad $P$ satisfies some technical conditions,
one can transfer the structure of a $P$-algebra by quasi-isomorphisms
of complexes. In particular, one can make the construction described
in the following lemma:
\begin{lemma}
Let $P$ be an operad of complexes, such that if we consider $P$ as an operad
just of $\mathbb{Z}$-graded vector spaces, it is free and generated by operations
in $\ge 2$ arguments. Let $A$ be an algebra over $P$, and let us choose
a splitting of $A$ considered as a $\mathbb{Z}$-graded space into the direct sum
$$A=H^*(A)\oplus V\oplus V[-1],\,\,\,(V[-1])^k:=V^{k-1}$$
endowed with a differential of the form
$d(a\oplus b\oplus c)=0\oplus 0\oplus b[-1]$.
Then there is a canonical structure of a $P$-algebra on the cohomology
space $H^*(A)$, and this algebra is homotopy equivalent to $A$.
\end{lemma}
We shall use it later (in 3.5) in combination with the fact that the
operad $\mathbb{Q}\otimes \mathsf{Chains}(C_d)$ is free as an operad of $\mathbb{Z}$-graded
vector spaces over $\mathbb{Q}$. This is evident because the action of $S_n$ on
$C_d(n)$ is free and the composition morphisms in $C_d$ are embeddings.
One can associate with any operad $P$ of complexes and, with any $P$-algebra
$A$, some differential graded Lie algebra (or more generally, a
$L_{\infty}$-algebra, see 3.1) $\mathsf{Def(A)}$. This Lie algebra is
defined canonically up to a quasi-isomorphism (the same as up to a homotopy).
It controls the deformations of $P$-algebra structure on $A$.
There are several equivalent constructions of $\mathsf{Def}(A)$ using
either resolutions of $A$ or resolutions of the operad $P$.
Morally, $\mathsf{Def}(A)$ is the Lie algebra of derivations in homotopy
sense of $A$. For example, if $P$ is an operad with zero differential then
$\mathsf{Def}(A)$ is quasi-isomorphic to the differential graded Lie algebra
of derivations of $\tilde{A}$ where $\tilde{A}$ is any free resolution of $A$.
The differential graded Lie algebras $\mathsf{Def}(A_1)$ and
$\mathsf{Def}(A_2)$ are quasi-isomorphic for homotopy equivalent
$P$-algebras $A_1$ and $A_2$.
\subsection{Hochschild complexes and deformation theory}
First of all, if $A$ a $d$-algebra then the shifted complex $A[d-1]$,
$$(A[d-1])^k:=A^{(d-1)+k}$$
carries a natural structure of $L_{\infty}$-algebra. It comes from a
homomorphism of operads in homotopy sense from the twisted by $[d-1]$
operad $\mathsf{Chains}(C_d)$ to the operad $\mathsf{Lie}$. In order to
construct such a homomorphism one can use fundamental chains of all
components of the Fulton-MacPherson operad (see 3.3.1), or deduce the
existence of a homomorphism from the results in 3.2.
Moreover, $A[d-1]$ maps as homotopy Lie algebra to $\mathsf{Def}(A)$, i.e.
$A[d-1]$ maps to ``inner derivations'' of $A$. These inner derivations
form a Lie ideal in $\mathsf{Def}(A)$ in homotopy sense.
\begin{claim}
The quotient homotopy Lie algebra $\mathsf{Def}(A)/A[d-1]$ is naturally
quasi-isomorphic to $\mathsf{Hoch}(A)[d]$.
\end{claim}
In the case when $d=0$ and the complex $A$ is concentrated in degree $0$,
the Lie algebra $\mathsf{Def}(A)$ is $\mathsf{End}(A)$, i.e. the Lie
algebra of linear transformations in the vector space $A$.
\begin{lemma}
The Hochschild complex of $0$-algebra $A$ is $A\oplus \mathsf{End}(A)$
(placed in degree $0$).
\end{lemma}
Here follows a sketch of the proof. First of all, the colored operad $SC_0$
is quasi-isomorphic to its zero-homology operad $H_0(SC_0)$ because all
connected components of spaces $(SC_0(n,m))_{n,m\ge 0}$ and of
$(C_1(n))_{n\ge 0}$ are contractible. By general philosophy (see 3.1)
this implies that we can replace $SC_0$ by $H_0(SC_0)$ in the definition
of the Hochschild complex given in 2.5. The $H_0$-version of a $1$-algebra
is an associative non-unital algebra, and the $H_0$-version of an action
is the following:
1) an action of an associative non-unital algebra $B$ on vector space $A$
(it comes from the generator of $\mathbb{Z}=H_0(SC_0(1,1)$),
2) a homomorphism from $B$ to $A$ of $B$-modules
(coming from the generator of $\mathbb{Z}=H_0(SC_0(1,0)$).
It is easy to see that to define an action as above is the same as to
define a homomorphism of non-unital associative algebras from $B$ to
$\mathsf{End}(A)\oplus A$. Thus, the Hochschild complex is (up to homotopy)
equal to $\mathsf{End}(A)\oplus A$.
Let us continue the explanation of the Claim 2 above for the case $d=0$.
As (homotopy) Lie algebra $\mathsf{Hoch}(A)$ coincides with the Lie algebra
of affine transformations on $A$. The homomorphism
$\mathsf{Def}(A){\longrightarrow} \mathsf{Hoch}(A)$ is a \textit{monomorphism}, but
in homotopy category every morphism of Lie algebras can be replaced
by an epimorphism! The Abelian Lie superalgebra $A[-1]$ is the ``kernel''
of this morphism. Let us show explicitly how all this works.
The Lie algebra $\mathsf{Def}(A)=(A)$ is quasi-isomorphic
to the following differential graded Lie algebra $\mathfrak{g}$: as $\mathbb{Z}$-graded space
it is $$\mathsf{End}(A)\oplus A\oplus A[-1],$$
i.e. the graded components of $\mathfrak{g}$ are
$\mathfrak{g}^0=\mathsf{End}(A)\oplus A,\,\,\,\mathfrak{g}^1=A,\,\,\,\mathfrak{g}^{\ne 0,1}=0$.
The nontrivial components of the Lie bracket on $\mathfrak{g}$ are
the usual bracket on $\mathsf{End}(A)$ and the action of $\mathsf{End}(A)$
on $A$ and on $A[-1]$. The only nontrivial component of the differential
on $\mathfrak{g}$ is the shifted by [1] identity map from $A$ to $A[-1]$.
The evident homomorphism $$\mathfrak{g}{\longrightarrow} \mathsf{End}(A)$$
is a homomorphism of differential graded Lie algebras, and
also a quasi-isomorphism.
There is a short exact sequence of dg-Lie algebras
$$0{\longrightarrow} A[-1]{\longrightarrow} \mathfrak{g}{\longrightarrow} \mathsf{End}(A)\oplus A{\longrightarrow} 0$$
In the case $d=1$ an analogous thing happens.
The deformation complex of an associative algebra $A$ is
the following \textit{subcomplex} of the shifted by [1] Hochschild complex:
$$\mathsf{Def}(A)^n:=\mathsf{Hom}_{\mathrm{vector\,spaces}}
(A^{\otimes(n+1)},A)\,\,\,{\rm for}\,\,n\ge 0; \,\,\,\,
\mathsf{Def}^{<0}(A):=0$$
The deformation complex is quasi-isomorphic to an $L_{\infty}$-algebra
$\mathfrak{g}$ which as $\mathbb{Z}$-graded vector space is
$$\mathsf{Def}(A)\oplus A\oplus A[1].$$
The Hochschild complex of $A$ is a \textit{quotient} complex of $\mathfrak{g}$
by the homotopy Lie ideal $A$.
\section{Formality of operads of chains, application to\\
deformation quantization}
\subsection{Quasi-isomorphisms of operads}
Operads themselves are algebras over a certain colored operad,
$\mathsf{Operads}$. This is quite obvious because an operad is just a
collection of vector spaces and polylinear maps between these spaces
satifying some identities. If we work in characteristic zero, it is
convenient to associate colors with all Young diagrams, i.e. with all
irreducible representations of all finite symmetric groups $S_n,\,\,n\ge 0$.
Analogously to the case of algebras, we can speak about
quasi-isomorphisms of operads of complexes.
\begin{definition}
A morphism $f:P_1{\longrightarrow} P_2$ between two operads of complexes is called
a quasi-isomorphism iff the maps of complexes $f(n):P_1(n){\longrightarrow} P_2(n)$
induce isomorphisms of cohomology groups for all $n$.
\end{definition}
Homotopy categories and deformation theories of algebras over
quasi-isomorphic operads are equivalent.
A typical example: the operad $\mathsf{Lie}$ is quasi-isomorphic to the
operad $L_{\infty}$ describing $L_{\infty}$-algebras.
Remind that a $L_{\infty}$-algebra $V$ is a complex of vector spaces
endowed with a coderivation $d_C$ of degree $(+1)$
of the cofree cocommutative coassociative
$\mathbb{Z}$-graded coalgebra without counit cogenerated by $V[1]$:
$$C:=\oplus_{n\ge 1} \bigl((V[1])^{\otimes n}\bigr)_{S_n}$$
such that the component of $d_C$ mapping $V$ to $V$ coincides
(up to a shift) with the differential $d_V$ of $V$.
Analogously, the operad $\mathsf{Assoc}$ is quasi-isomorphic to the operad
$\mathsf{Chains}(C_1)$, and also to the operad $A_{\infty}$ responsible for
$A_{\infty}$-algebras.
\subsection{Formality of chain operads}
\begin{theorem}
The operad $\mathsf{Chains}(C_d)\otimes \mathbb{R}$ of complexes of real vector
spaces is quasi-isomorphic to its cohomology operad endowed with zero
differential.
\end{theorem}
In general, differential graded algebras quasi-isomorphic to their
cohomology endowed with zero differential, are called formal.
The classical example is the de Rham complex of a compact K\"ahler
manifold. The result of Deligne-Griffiths-Morgan-Sullivan (see [DGMS])
is that this algebra is formal as differential graded commutative
associative algebra. Our theorem says that $\mathsf{Chains}(C_d)\otimes \mathbb{R}$
is formal as an algebra over the colored operad $\mathsf{Operads}$.
The story of this theorem is truly complicated. It was stated in the
preprint of Getzler and Jones originally for the cases $d=1$ and $d=2$.
The case $d=1$ is trivial (as well as $d=0$). The authors referred to me
for the claim that $\mathsf{Chains}(C_d)\otimes \mathbb{R}$ is not formal for
$d\ge 3$. Later it became clear that the proof in the Getzler-Jones preprint
is not correct. Personally, I thought for several years that even the fact
is not true, and made several times calculations demonstrating the
non-formality of $\mathsf{Chains}(C_2)\otimes \mathbb{R}$. The story began to
stir again when D.~Tamarkin posted on the net a new proof of the formality
theorem in deformation quantization. The basic intermediate result
in Tamarkin's approach was the formality of $\mathsf{Chains}(C_2)\otimes \mathbb{R}$
for which he claimed a new (with respect to [GJ]) proof.
Unfortunately, there were several mistakes in Tamarkin's proof at that time
as well, until a new corrected version was posted (see [T1]) which did
not use Deligne's conjecture directly. In [T1] is proven that $H_*(C_2)$
acts on the Hochschild complex of any algebra $A$.
Also Tamarkin found a proof that $\mathsf{Chains}(C_2)\otimes \mathbb{C}$
is quasi-isomorphic to $H_*(C_2)\otimes \mathbb{C}$ (see [T2]). Deligne's conjecture
follow from the combination of two results of Tamarkin, but unfortunately
not in a purely topological/combinatorial way.
In the meantime I found a flaw in my calculations which ``show''
non-formality of $\mathsf{Chains}(C_2)\otimes \mathbb{R}$, and found a new proof
of the formality of $\mathsf{Chains}(C_d)\otimes \mathbb{R}$ valid for all $d$.
The quasi-isomorphism which I constructed differs essentially from the
one constructed by Tamarkin in [T2]. It seems that it gives really
different deformation quantizations of Poisson manifolds. The question of
differences between quantizations is addressed in the third part of
this article.
I finish this subsection by a description of the homology operads of
topological operads $C_d$.
\begin{theorem}
An algebra over $H_*(C_d)$ is
\noindent 1) a complex in the case $d=0$,
\noindent 2) a differential graded associative algebra in the case $d=1$,
\noindent 3) a differential graded ``twisted'' Poisson algebra with
the commutative associative product of degree $0$ and
with the Lie bracket of degree $(1-d)$ in the case of odd $d\ge 3$,
\noindent 4) a differential graded ``twisted'' Gerstenhaber algebra
with the commutative associative product of degree $0$ and with the
Lie bracket of degree $(1-d)$ in the case of even $d\ge 2$.
In cases 3) and 4) the Lie bracket satisfies the Leibniz rule with respect
to the product.
\end{theorem}
The ``twisting'' in cases 3) and 4) means just that the commutator
has the usual $\mathbb{Z}/2\mathbb{Z}$-grading, but a \textit{different} $\mathbb{Z}$-grading.
The total rank of the homology group $H_*(C_d(n))$ is $n!$ for all $d\ge
1$. The rank of the top degree homology group of $C_d(n))$ is $(n-1)!$
for any $d\ge 2$ and $n\ge 1$. As a representation of $S_n$ this
homology group $H_{(n-1)(d-1)}(C_d(n))$ is isomorphic (up to tensoring
by the sign representation of $S_n$ for even $n$) to $\mathsf{Lie}(n)$,
the $n$-th space of the Lie operad.
The conclusion from two theorems in this section is that $d$-algebras for
$d\ge 2$ are essentially the same as twisted Poisson or Gerstenhaber
algebras (depending on the parity of $d$).
\subsection{Sketch of the proof of the formality of chain
operads}
The proof presented here is quite technical and it is not really essential
for the rest of this paper. The reader can skip it and go directly to
section 3.4.
\subsubsection{Fulton-MacPherson compactification}
Let us fix dimension $d\ge 1$. There is a modification $FM_d$ of
the topological operad $C_d$ which is more convenient to work with. The
idea to consider operad $FM_d$ was proposed by several people,
in particular by Getzler and Jones in [GJ]. The letters ``FM'' stand
for Fulton and MacPherson, who introduced closely related constructions
in the realm of algebraic geometry (see [FM]).
I shall describe now the operad $FM_d$.
For $n\ge 2$ denote by ${\tilde {C}}_d(n)$ the quotient space of
the configuration space of $n$ points in $\mathbb{R}^d$
$$\mathsf{Conf}_n(\mathbb{R}^d):=\{(x_1,\dots,x_n)\in (\mathbb{R}^d)^n|
\,x_i\ne x_j \,\,\,{\mathrm{for\,\,\,any}}\,\,\,i\ne j\}$$
modulo the action of the group $G_d=\{x\mapsto \lambda x +v|\, \lambda \in
\mathbb{R}_{>0},\,\,v\in \mathbb{R}^d\}$. The space ${\tilde {C}}_d(n)$ is a smooth manifold
of dimension $(nd -d-1)$. For $n=2$, the space ${\tilde {C}}_d(n)$ coincides with the
$(d-1)$-dimensional sphere $S^{d-1}$. There is an obvious free action of
$S_n$ on ${\tilde {C}}_d(n)$. We define the spaces ${\tilde {C}}_d(0)$ and ${\tilde {C}}_d(1)$ to be
empty. The collection of spaces ${\tilde {C}}_d(n)$ does not form an operad because
there is no identity element, and compositions are not defined.
The components of the operad $FM_d$ are
1) $FM_d(0):=\emptyset$,
2) $FM_d(1)=$point,
3) $FM_d(2)={\tilde {C}}_d(2)=S^{d-1}$,
4) for $n\ge 3$ the space $FM_d(n)$ is a manifold with corners, its
interior is $C'_d(n)$, and all boundary strata are certain products
of copies of ${\tilde {C}}_d(n')$ for $n'<n$.
A manifold with corners looks locally as a product of manifolds with
boundary. Any manifold with corners is automatically a \textit{topological}
manifold, although its \textit{smooth} structure is not the one of
a differentiable manifold with boundary.
Intuitively, if a configuration of $n$ points in $\mathbb{R}^d$ moves in such a
way that several points (or groups of points) become at the limit close to
each other, we use a microscope with a very large magnification
(apply a large element of the group $G_d$) in order to see in details
the shape of configurations of points in clusters.
One of the rigorous definitions of $FM_d(n)$ is the following:
\begin{definition}
For $n\ge 2$, the manifold with corners $FM_d(n)$ is the closure
of the image of ${\tilde {C}}_d(n)$ in the compact manifold
$\left(S^{d-1}\right)^{n(n-1)/2}$ under the map
$$G_d\cdot (x_1,\dots,x_n)\mapsto
\left({x_j-x_i\over |x_j-x_i|}\right)_{1\le i<j\le n}$$
\end{definition}
Set-theoretically, the operad $FM_d$ is the same as the free operad
generated by the collection of sets $({\tilde {C}}_d(n))_{n\ge 0}$ endowed with
$S_n$-actions as above.
It is possible to define another topological operad $FM'_d$, and
two homomorphisms of operads
$$f_1:C_d{\longrightarrow} FM'_d,\,\,\, f_2:FM_d{\longrightarrow} FM'_d$$
such that for all $n\ge 0$ maps
$$f_1(n):C_d(n){\longrightarrow} FM'_d(n),\,\,\, f_2(n):FM_d(n){\longrightarrow} FM'_d(n)$$
are homotopy equivalences. In a sense, the spaces $FM'_d(n)$ parametrize
configurations of small disks in the standard disk, together
with a class of ``degenerate'' configurations in which some
(or all) disks are infinitely small.
We leave as an exercise to the reader to give a complete definition of $FM'_d$.
Applying the functor $\mathsf{Chains}$ we get two quasi-isomorphisms
of operads of complexes
$$\mathsf{Chains}(C_d){\longrightarrow} \mathsf{Chains}(FM'_d)$$
$$\mathsf{Chains}(FM'_d){\longleftarrow} \mathsf{Chains}(FM_d)$$
\subsubsection{A chain of quasi-isomorphisms}
For any $d\ge 2$ I shall define several operads of complexes
and construct a chain of quasi-isomorphisms between them.
The total diagram is the following:
$$\mathsf{Chains}(FM_d) {\longleftarrow} \mathsf{SemiAlgChains}(FM_d)$$
\begin{equation}
\mathsf{SemiAlgChains}(FM_d)\otimes \mathbb{R} {\longrightarrow} \mathsf{Graphs}_d {\hat{\otimes}}\mathbb{R}
\end{equation}
$$\mathsf{Graphs}_d {\longleftarrow} \mathsf{Forests}_d{\longleftarrow} H_*(\mathsf{Forests)}H_*(C_d)$$
I shall now explain the first line. It is really technical, and is
introduced only to circumvent some difficulties with integrals which appear
in the second line (see the next section).
The operad $\mathsf{SemiAlgChains}(FM_d)$ is the suboperad of
$\mathsf{Chains}(FM_d)$ consisting of combinations of maps
$[0,1]^k{\longrightarrow} FM_d(n)$ whose graphs are real semi-algebraic sets. This is
well defined because the space $FM_d(n)$ can be described in terms of
algebraic equations and inequalities. It is a closed semi-algebraic subset
in the product of $\frac{1}{2}n(n-1)$ copies of $S^{d-1}$. The natural
inclusion of semi-algebraic chains into all continuous chains is
a quasi-isomorphism of operads.
\subsubsection{Admissible graphs and corresponding differential forms}
\medskip
\begin{definition}
An admissible graph with parameters $(n,m,k)$ (for $n\ge 1$ and $m\ge 0$)
is a finite graph ${\Gamma}$ such that
\noindent 1) it has no multiple edges,
\noindent 2) it contains no simple loops (edges connecting a vertex
with itself),
\noindent 3) it contains $n+m$ vertices, numbered from $1$ to $n+m$,
\noindent 4) it contains $k$ edges, numbered from $1$ to $k$,
\noindent 5) any vertex can be connected by a path with a vertex whose
index is in $\{1,\dots,n\}$,
\noindent 6) any vertex with index in $\{n+1,\dots,n+m\}$ has valency
(i.e. degree) $\ge3$.
greater than or equal to $3$.
\noindent 7) for every edge $E$ of ${\Gamma}$ we choose an orientation of this edge,
i.e. we order the 2-element set of vertices to which $E$ is attached.
\noindent 8) if $n=1$ then the graph consists of just one vertex and has
no edges, its parameters are $(1,0,0)$.
\end{definition}
Notice that although we endowed edges with orientations in 7),
we treat in 1)-6) the graph ${\Gamma}$ as an \textit{unoriented} graph.
The structure of an admissible graph is completely determined by the
attachment map
$$\{1,\dots,k\}{\longrightarrow} \{(i,j)|1\le i,j,\le n+m,\,\,i\ne j\}$$
from the set of edges to the set of ordered pairs of distinct vertices.
\begin{definition}
Let ${\Gamma}$ be an admissible graph. We define $\omega_{{\Gamma}}$ to be the
differential form on $FM_d(n)$ given by the formula
\begin{equation}
\omega_{{\Gamma}}:=(\pi_1)_*\circ \pi_2^*\left(\bigwedge_{{\mathrm{edges\,\,of\,\,}}
{\Gamma}} \mathrm{Vol}_{S^{d-1}}\right)
\end{equation}
where
$$\pi_1:FM_d(n+m){\longrightarrow} FM_d(n)$$ is the natural map, defined by forgetting
the last $m$ points in the configuration of $(n+m)$ points in $\mathbb{R}^d$,
$$\pi_2:FM_d(n+m){\longrightarrow} \left( FM_d(2)\right)^k$$
is the product of forgetting maps $FM_d(n+m){\longrightarrow} FM_d(2)=S^{d-1}$
associated with the edges of ${\Gamma}$ (i.e. with ordered pairs of indices in
$\{1,\dots,n+m\}$),
$$Vol_{S^{d-1}}\in \Omega^{d-1}(S^{d-1})$$ denotes the volume form on
$S^{d-1}$ invariant under the action of the rotation group $SO(d,\mathbb{R})$
and normalized such that the total volume
$\int_{S^{d-1}}\mathrm{Vol}_{S^{d-1}}$ is $1$.
\end{definition}
The degree of the form $\omega_{{\Gamma}}$ is equal to
$$(d-1)k-dm=\mathrm{dim}\left(FM_d(2)^k\right)-\left(\mathrm{dim}(FM_d(n+m)-
\mathrm{dim}(FM_d(n))\right).$$
If one changes orientations of edges, or the enumeration of edges,
or the enumeration of vertices with indices from $n+1$ to $n+m$, one
obtains the same form up to a sign. We leave as an easy exercise to the
reader to write an explicit formula for this sign.
\begin{definition}
For every $n\ge 1$ we define $\mathsf{Graphs}_d(n)$ to be the
$\mathbb{Z}$-graded vector space of all $\mathbb{Q}$-valued functions on
the set of equivalence classes of admissible graphs with parameters
$(n,m,k)$, $m$ and $k$ arbitrary, such that if we change
the enumeration of edges, or the enumeration of vertices with
indices from $n+1$ to $n+m$, then the value of the function will be
multiplied by an appropriate sign as explained above.
We define $\mathbb{Z}$-grading of a function concentrated on one given
equivalence class $[{\Gamma}]$ with parameters $(n,m,k)$ as
$\left(dm-(d-1)k\right)$.
\end{definition}
Unfortunately, the forms $\omega_{\Gamma}$ are not $C^{\infty}$-forms on the
boundary of $FM_d(n)$. Still, for any ${\Gamma}$ the integral of $\omega_{{\Gamma}}$
over any semi-algebraic chain is absolutely convergent because the
calculation of this integral reduces to the calculation the volume form
over a compact semi-algebraic chain of the top degree in the product
of spheres. The total volume is finite because the multiplicity of a
semi-algebraic map is bounded. Thus the integral of the volume form is
convergent.
{}From this follows that the form $\omega_{\Gamma}$ gives a well-defined functional
on $\mathsf{SemiAlgChains}(FM_d)$. Turning things around, we can say
that any semi-algebraic chain gives a functional on the set
of equivalence classes of admissible graphs, i.e. an element
of the real-valued version of $\mathsf{Graphs}_d(n)$, the completed tensor
product $\mathbb{R}\hat\otimes \mathsf{Graphs}_d(n)$. The difference between the
completed and the usual tensor product by $\mathbb{R}$ will eventually disappear
because we shall meet the operad $H_*(FM_d)$ whose components
are \textit{finite-dimensional}.
\begin{lemma}
For any ${\Gamma}$ the form $d\omega_{\Gamma}$ (considered as a functional on
semi-algebraic chains) is equal to the sum with appropriate signs of
forms $\omega_{{\Gamma}'}$ where the admissible graph ${\Gamma}'$ is obtained
from ${\Gamma}$ by contraction of one edge.
\end{lemma}
This lemma was proven in [K1] (lemma in 6.6) for the case $d=2$ and
in [K2] (lemma 2.1) for the case $d\ge 3$.
Thus, the graded space $\mathsf{Graphs}_d(n)$ of functions on graphs
carries a naturally defined differential, and forms a complex of vector
spaces over $\mathbb{Q}$. Moreover, the restrictions of the forms $\omega_{\Gamma}$ to
irreducible components of the boundary of the manifolds $FM_d(n)$ are
finite linear combinations of products of analogous forms for simpler graphs.
This means that $\mathsf{Graphs}_d$ is an operad of complexes,
and the integration defines a homomorphism of operads of complexes of
real vector spaces
$$\mathbb{R} {\otimes}\mathsf{SemiAlgChains}(FM_d){\longrightarrow}
\mathbb{R} {\hat\otimes}\mathsf{Graphs}_d $$
It is not obvious that this arrow is a quasi-isomorphism.
This follows from an explicit calculation of the cohomology operad of the
graph operad which we perform in the next subsection.
\subsubsection{Forests and Tree complexes}
\medskip
\begin{definition}
An admissible graph is a forest iff it contains no nontrivial closed paths.
\end{definition}
It is easy to see that for non-forest graph ${\Gamma}$ the differential
$d\omega_{{\Gamma}}$ is a linear combination of forms $\omega_{{\Gamma}'}$ for
non-forest graphs ${\Gamma}'$. Also, the restriction of $\omega_{{\Gamma}}$ to
irreducible components of the boundary of $FM_d(n)$ is a linear
combination of products $\omega_{{\Gamma}_{1}}\times\omega_{{\Gamma}_{2}}$
where at least one of smaller graphs ${\Gamma}_1,{\Gamma}_2$ is not a forest.
All this implies that the following definition gives an operad:
\begin{definition}
The Operad $\mathsf{Forests}_d$ is a suboperad of $\mathsf{Graphs}_d$
consisting of functions vanishing on all non-forest graphs.
\end{definition}
A simple spectral sequence shows that the embedding
$$\mathsf{Forests}_d{\longrightarrow} \mathsf{Graphs}_d$$
is a quasi-isomorphism. We shall not write it here explicitly, only
mention the main idea of the calculation. Any graph which is not a forest
contains a nonempty maximal subgraph ${\Gamma}_{\mathrm{core}}$ such that
the valency (in ${\Gamma}_{\mathrm{core}}$) of each vertex of ${\Gamma}_{\mathrm{core}}$
is at least $2$. The graph ${\Gamma}$ is obtained from ${\Gamma}_{\mathrm{core}}$ by
attaching several trees to vertices of ${\Gamma}_{\mathrm{core}}$, and also a forest
not connected with ${\Gamma}_{\mathrm{core}}$. The desired result follows from
the vanishing of the first term of the spectral sequence associated with
the filtration by the number of vertices in ${\Gamma}_{\mathrm{core}}$ on
the quotient complex ${\mathsf{Graphs}}_d(n)/{\mathrm{Forests}}_d(n)$.
Any admissible graph ${\Gamma}$ with parameters $(n,m,k)$ defines
a partition of $\{1,\dots,n\}$ into pieces corresponding to
connected components of ${\Gamma}$. All the graphs ${\Gamma}'$ which appear in the
decomposition of $d\omega_{{\Gamma}}$ (as in the lemma in 3.3.3) give the same
partition of $\{1,\dots,n\}$ as ${\Gamma}$. This implies that the forest
complex ${\mathsf{Forests}}_d(n)$ splits naturally for any $n$ into
a direct sum of subcomplexes corresponding to partitions
(i.e. equivalence relations) of the set $\{1,\dots,n\}$.
The subcomplex of $\mathsf{Forests}_d(n)$ associated with a partition
is the tensor product of tree complexes over pieces of this partition,
where \textit{trees} are non-empty connected forests, as usual.
We show an example of the tree complex in the case of $3$ vertices.
There are $4$ admissible trees with the parameter $n$ equal to $3$
(up to changing the enumeration of edges
and orientations of edges):
\begin{center}
\scalebox{1}{\includegraphics{maximf1.eps}}
\end{center}
The formula for the de Rham differential of forms $\omega_{{\Gamma}_i}$ is
$$d\omega_{{\Gamma}_1}=d\omega_{{\Gamma}_2}=d\omega_{{\Gamma}_3}=0,\,\,\,d\omega_{{\Gamma}_4}
=\pm\omega_{{\Gamma}_1}\pm\omega_{{\Gamma}_2}\pm\omega_{{\Gamma}_3}\,\,\,.$$
The tree complex is the \textit{dual} complex to the complex spanned by
the differential forms $\omega_{{\Gamma}}$ for trees ${\Gamma}$. It is easy to see that
the tree complex has the non-zero cohomology space only in the lowest
degree. This implies that there is a natural map from the cohomology of the
tree complex to the tree complex. Thus, we get a natural morphism from
the cohomology of the forest complex to the forest complex.
One can check that this map is a morphism of operads, and we get a
quasi-isomorphism $$H^*(\mathsf{Forests}_d){\longrightarrow} \mathsf{Forests}_d$$
Comparing the cohomology of the forest complex with known results
on the cohomology of configuration spaces, we see that
$$H^*(\mathsf{Forests}_d)=H_*(C_d)=H_*(FM_d)\,\,\,.$$
Because we proved already that
$H^*(\mathsf{Graphs}_d)=H^*(\mathsf{Forests}_d)$,
we constructed a chain of quasi-isomorphisms as promised.
\subsection{Application to deformation quantization}
The theorems in 3.2 show that any Gerstenhaber algebra, i.e.
an algebra over $H_*(C_d)$, can be canonically endowed with a $d$-algebra
structure. We prove in this section the result of Tamarkin:
\begin{theorem}
Let $A:=\mathbb{R}[x_1,\dots,x_n]$ be the algebra of polynomials
considered just as an associative algebra.
Then the Hochschild complex $\mathsf{Hoch}(A)$ is quasi-isomorphic
as $2$-algebra to its cohomology
$$B:=H^*(\mathsf{Hoch}(A))={\rm space\,\,of \,\,polynomial\,\,
polyvector\,\,fields\,\,on\,\,\,}\mathbb{R}^n$$
considered as a Gerstenhaber algebra, hence a $2$-algebra.
\end{theorem}
As was mentioned in the introduction, this theorem implies
the formality theorem from [K2].
It is well-known (Hochschild-Kostant-Rosenberg theorem) that
the space of polyvector fields is equal to the cohomology of
$\mathsf{Hoch}(A)$, and the cup-product and the Lie bracket on
$B=H^*(\mathsf{Hoch}(A))$ are the usual cup-product and the
Schouten-Nijenhuis bracket on polyvector fields, respectively.
{}From the general formalism of deformation theory it follows that
if $\mathsf{Hoch}(A)$ is \textit{not} quasi-isomorphic to $B$ then
there will be a \textit{non-zero} obstruction
element $\gamma_{\mathrm{obstr}}$ in the first cohomology group
of the deformation complex $\mathsf{Def}(B)$ of $2$-algebra $B$.
Moreover, there exists an $\mathsf{Aff}(\mathbb{R}^n)$-invariant splitting of
$\mathsf{Hoch}(A)$ into the direct sum of $B$ and a splitted contractible
complex (see arguments in [K1], 4.6.1.1). By the lemma in 2.6, this
splitting induces a structure of $2$-algebra on $B$, and this structure
is $\mathsf{Aff}(\mathbb{R}^n)$-invariant. It implies that
$\gamma_{\mathrm{obstr}}\in H^1(\mathsf{Def}(B))$ is also
$\mathsf{Aff}(\mathbb{R}^n)$-invariant.
Let us calculate the cohomology of the deformation complex of the
$2$-algebra $B$. As a first approximation, we calculate the Hochschild
cohomology of $B$:
\begin{theorem}
The Hochschild complex of $B$ is quasi-isomorphic to $\mathbb{R}^1$
placed in degree $0$.
\end{theorem}
In order to prove it we need an additional result concerning the
Hochschild cohomology of $d$-algebras with trivial Lie bracket.
By results from 3.2 we can speak about algebras over $H_*(C_d)$ instead
of $d$-algebras.
\begin{lemma}
Let $d\ge 2$ be an integer. Consider the algebra of polynomials in a finite
number of $\mathbb{Z}$-graded variables
$$\mathbb{R}[x_1,\dots,x_N],\,\,\,deg(x_i)=d_i\in \mathbb{Z}$$
as an algebra over the operad $H_*(C_d)$, endowing it with zero differential
and with the vanishing Lie bracket. Then the Hochschild cohomology of
this algebra is, as $\mathbb{Z}$-graded vector space, the same as the algebra of
polynomials in the doubled set of variables
$$\mathbb{R}[x_1,\dots,x_N,y_1,\dots,y_N],\,\,\,deg(x_i)=d_i,\,\,deg(y_i)=d-d_i$$
\end{lemma}
The statement of this lemma is similar to the classical
Hochschild-Kostant-Rosenberg calculation of the Hochschild cohomology
of the algebra of polynomials considered as an associative algebra
(i.e. as $1$-algebra). We shall not give here the proof of this lemma.
In general, if $\mathcal{O}(X)$ is the algebra of functions on a smooth
$\mathbb{Z}$-graded algebraic super-manifold $X$, then the Hochschild
cohomology of $\mathcal{O}(X)$ considered as a $d$-algebra, coincides with
the algebra of functions on the total space of the twisted by $[d]$
cotangent bundle to $X$:
$$H^*(\mathsf{Hoch}(\mathcal{O}(X)))=\mathcal{O}(T^*[d] X)\,\,\,.$$
Applying the above lemma to the following $H_*(C_2)$-algebra with
\textit{vanishing} Lie bracket
$$B_0:=\mathbb{R}[x_1,\dots,x_n,\xi_1,\dots,\xi_n],\,\,\deg(x_i)=0,\,
\deg(\xi_i)=+1$$
we get the Hochschild cohomology
$$C_0:=\mathbb{R}[(x_i),(\xi_i),(\eta_i),(y_i)]_{1\le i\le n},$$
$$\deg(x_i)=0,\, \deg(\xi_i)=\deg(\eta_i)=
+1,\,\deg(y_i)=+2\,\,\,.$$
Our 2-algebra $B$ is obtained from $B_0$ by switching on the Lie bracket:
$$[x_i,x_j]=[\xi_i,\xi_j]=0,\,\,\,[x_i,\xi_j]=\delta_{ij}\,\,.$$
It is easy to see that this produces the following differential on $C_0$:
$$d_1:=\sum_{i=1}^n \bigl(\eta_i{\partial \over \partial x_i}
+y_i{\partial \over \partial \xi_i}\bigr)$$
The cohomology of the complex $(C_0,d_1)$ is equal to the de Rham
cohomology of $\mathbb{R}^n$, i.e. to $\mathbb{R}$ placed in degree $0$. What we
calculated is only the first term in a spectral sequence, but it is clear
that higher differentials are zero because there is no space for them.
This proves our theorem on the Hochschild cohomology of
$B=\mathsf{Hoch}[x_1,\dots,x_n]$:
\begin{equation}
H^*(\mathsf{Hoch}(\mathsf{Hoch}(\mathbb{R}[x_1,\dots,x_n])))=\mathbb{R}\,\,\,.
\end{equation}
Now we are ready to calculate the first cohomology of the deformation
complex of $B$. Remind (see 2.7) that there is a short exact sequence
$$0{\longrightarrow} B[1]{\longrightarrow} \mathsf{Def}(B){\longrightarrow} \mathsf{Hoch}(B)[2]{\longrightarrow} 0\,\,\,.$$
Passing from this exact sequence to the level of the cohomology
we get a long exact sequence
$$\cdots{\longrightarrow} H^{i+1}(\mathsf{Hoch}(B)){\longrightarrow} B^{i+1}{\longrightarrow}
H^i(\mathsf{Def}(B)){\longrightarrow} H^{i+2}(\mathsf{Hoch}(B)){\longrightarrow} \cdots$$
which for $i=1$ says that
$$H^1(\mathsf{Def}(B))=B^2={\Gamma}(\mathbb{R}^d,\wedge^2 T)=
\{ {\rm \,polynomial \,\,bivector\,\,fields\,\,on\,\,}\mathbb{R}^d\}$$
Then the argument goes as was sketched in the introduction.
The explicit quasi-isomorphism between the space of polyvector fields
and the Hochschild complex of the associative algebra
$A=\mathbb{R}[x_1,\dots,x_n]$ can be made invariant under the action of the group of
affine transformations. There is no non-zero $\mathsf{Aff}(\mathbb{R}^n)$-invariant
bi-vector fields on $\mathbb{R}^n$. Thus the first non-trivial obstruction
$\gamma_{\mathrm{obstr}}$ to the existence of a quasi-isomorphism between
$B$ and $\mathsf{Hoch}(A)$ cannot exist, and $B$ is quasi-isomorphic to
$\mathsf{Hoch}(A)$.
\section{Grothendieck-Teichm\"uller group action on
quantizations}
\subsection{Integrals in deformation quantization}
There are several situations in deformation quantization where coefficients
in formulas are given by explicit integrals.
A) In Drinfeld's study of classical Knizhnik-Zamoldchi\-kov equations
(see [D]), a formal series in two non-commuting variables
appears, called an \textit{associator}. The coefficients in the explicit
formula for the associator are iterated integrals
\begin{equation}
I_{{\epsilon}_1,\dots,{\epsilon}_n}:=\int_{0<t_1<\dots<t_n<1} \omega_{{\epsilon}_1}(t_1)
\wedge\dots\wedge\omega_{{\epsilon}_n}(t_n)
\end{equation}
where ${\epsilon}_i\in\{0,1\},\,\,\,{\epsilon}_0=1,{\epsilon}_n=0,\,\,\,
\omega_0(t)=dt/t,\,\,\,\omega_1(t)=dt/(1-t)$.
Among these numbers there are values of the Riemann zeta-function at positive
integers, $\zeta(n)=I_{1,0,0,\dots,0}$ for a $n$-dimensional integral.
In general, the integrals $I_{{\epsilon}_1,\dots,{\epsilon}_n}$ can be identified with
so called \textit{multiple zeta-values} (see [Z]).
B) The same class of numbers appears in the Etingof-Kazhdan quantization of
Poisson-Lie algebras because it was based on Drinfeld's work (see [EK]).
C) Tamarkin uses the Drinfeld associator and Etingof-Kazhdan results
from [EK] in his proof of the formality of $\mathsf{Chains}(C_2)\otimes \mathbb{C}$
(see [T2]).
D) Tamarkin uses also the Drinfeld associator in the new proof of the
formality theorem in deformation quantization (see [T1]).
E) In my construction of quantization of Poisson manifolds (see [K1])
other types of integrals were used. These integrals are real-valued and
are expressed by:
\begin{equation}
\int_{U_{n,m}} \bigwedge_{k=1}^{m+2n-2}\alpha(z_{i_k},z_{j_k}),\,\,m\ge 0,
n\ge 1
\end{equation}
where the domain of integration $U_{n,m}$ is
\begin{eqnarray*}
&& \{(z_1,\dots,z_{n+m})\in\mathbb{C}^{n+m}|\,
z_1,\dots,z_m\in \mathbb{R}, z_1<\dots< z_m;\\
&&\quad \Im(z_{m+1}),\dots,\Im(z_{m+n})>0;z_i\ne z_j;
z_{m+n}=\sqrt{-1}\}
\end{eqnarray*}
The form $\alpha(z,w)$ is
\begin{equation}
{1\over 4 \pi i } d\log\left({(z-w)(z-{\overline w})
\over ({\overline z}-w)({\overline z}-{\overline w})}\right)\,.
\end{equation}
The range of indices $i_k\ne j_k$ is $\{1,\dots, n+m\}$ for all
$k\in\{1,\dots,m+2n-2\}$.
F) Recently I realized that one can modify the above formulas,
replacing $\alpha$ by
\begin{equation}
\alpha_{new}(z,w):= d\log\left({(z-w)\over ({\overline z}-w)}\right)
\end{equation}
and dividing the integral by $(2\pi i)^n$.
In this way one gets complex-valued integrals, and all identities
proven in [K2] remain true.
G) The last example is the use of integrals in the proof of formality of
$\mathsf{Chains}(C_2)\otimes \mathbb{R}$ presented in section 3.3.
I claim that all these integrals are closely related, and their use is
probably unavoidable.
\subsection{Torsors}
In all the above situations A)-G), we are constructing
\textit{identifications} (up to a homotopy)
between certain pairs of algebraic structures. For example, in
deformation quantization we construct an isomorphism in the homotopy
category of Lie algebras between the shifted by $[1]$
Hochschild complex of $\mathbb{R}[x_1,\dots,x_n]$, and
the graded Lie algebra of polyvector fields on $\mathbb{R}^n$.
Let $C$ be any category and $\mathcal{E},\mathcal{F}$ be two isomorphic objects
in this category. The set of isomorphisms $$\mathsf{Iso}(\mathcal{E},\mathcal{F})$$
is a non-empty set on which the group $\mathsf{Aut}(\mathcal{E})$ acts simply
transitively. The group $Aut(\mathcal{F})$ acts also simply transitively on it,
and the two actions commute. One can encode all these structures in a
single map
\begin{equation}
\mathsf{Iso}(\mathcal{E},\mathcal{F})^3{\longrightarrow}
\mathsf{Iso}(\mathcal{E},\mathcal{F}),\,\,\,(a,b,c)\mapsto a\circ b^{-1}\circ c
\end{equation}
\begin{definition}
A torsor is a non-empty set $X$ endowed with a map
$X\times X\times X{\longrightarrow} X$ satisfying the same identities as
maps $(a,b,c){\longrightarrow} a b^{-1} c $ in groups.
\end{definition}
A torsor is the same as a principal homogeneous space over a group.
One can give a more transparent definition of essentially the same
structure:
\begin{definition}
A torsor is a category $C$ with only two objects $\mathsf{Ob}_1$ and
$\mathsf{Ob}_2$, such that all the morphisms in $C$ are invertible and the
objects $\mathsf{Ob}_1$ and $\mathsf{Ob}_2$ are equivalent.
\end{definition}
If $X$ is a torsor (by Definition (18)) then one has \textit{two}
groups acting simply transitively on $X$. Any element $x\in X$
gives an identification between these two groups.
\subsubsection{Pro-algebraic torsors}
In each of the cases listed in 4.1 in order to construct
an isomorphism one is brought to solve an infinite system of quadratic
equations with integer coefficients. For example, in deformation
quantization we need to choose weights for all finite graphs in order to
get a quasi-isomorphism between two homotopy Lie algebras. This
implies that the torsor of isomorphisms should be considered
as an infinite-dimensional algebraic variety over $\mathbb{Q}$, not only
as a set. We denote the torsor of isomorphisms considered
as an algebraic variety by
$${\underline {\mathsf{Iso}}}(\mathcal{E},\mathcal{F})$$
Each of the two groups acting on the torsor is a projective limit
of finite-dimensional affine algebraic groups over $\mathbb{Q}$.
The algebra of functions $\mathcal{O}(T)$ on a pro-algebraic torsor $T$
is a generalization of a Hopf algebra. Namely, it is a commutative
associative unital algebra over $\mathbb{Q}$ together with the structure map
\begin{equation}
\mathcal{O}(T)\mapsto \mathcal{O}(T)\otimes \mathcal{O}(T)\otimes \mathcal{O}(T)
\end{equation}
which is a homomorphism of algebras, and satisfies relations dual to
the defining relations for set-theoretic torsors.
We call this map the \textit{triple coproduct} in $\mathcal{O}(T)$.
The integrals in situations A)-G) from 4.1 give solutions to systems
of quadratic equations in complex numbers, i.e. they give a homomorphism
of algebras over $\mathbb{Q}$ $$\mathcal{O}(T){\longrightarrow} \mathbb{C}\,\,\,$$
where the pro-algebraic torsor $T$ depends on the concrete problem
which we consider. In the next subsection we are going to describe a
countable-dimensional commutative algebra $P$ over $\mathbb{Q}$ such that the
homomorphism from $\mathcal{O}(T)$ to $\mathbb{C}$ is the composition of homomorphisms
$\mathcal{O}(T){\longrightarrow} P$ and $P{\longrightarrow} \mathbb{C}$.
\subsection{Periods, motivic Galois group, motives }
Periods are integrals of algebraic differential forms with algebraic
coefficients. The following numbers are periods:
$$ 1,\,\,\sqrt{2},\ i=\sqrt{-1},\ \pi,\ \log(2),$$
$$\zeta(3)=\int_{0<t_1<t_2<t_3<1}
{dt_1\over 1-t_1} {dt_2\over t_2}{dt_3\over t_3},\
\mathrm{elliptic\,\, integral}\ \int_1^2\sqrt{x^3+1}dx,\dots$$
We shall need a more precise definition of periods.
Let $X$ be a smooth
algebraic variety of dimension $d$
defined over $\mathbb{Q}$, $D\subset X$ be a divisor with normal crossings
(i.e. locally $D$ looks like a collection of coordinate hypersurfaces),
$\omega \in \Omega^d(X)$ be an algebraic differential
form on $X$ of top degree ($\omega$ is automatically closed),
and $\gamma\in H_d(X(\mathbb{C}),D(\mathbb{C});\mathbb{Q})$ be a (homology class of a)
singular chain on the complex manifold $X(\mathbb{C})$ with boundary on
the divisor $D(\mathbb{C})$. With these data one associates the integral
$\int_\gamma \omega \in \mathbb{C}$. We say that this number
is the period of the quadruple $(X,D,\omega,\gamma)$.
One can always reduce convergent integrals of algebraic forms over
semi-algebraic sets defined over the field algebraic numbers $\overline{\mathbb{Q}}$
to the form as above, using the functor of the restriction of scalars
to $\mathbb{Q}$ and the resolution of singularities in characteristic zero.
Usual tools for proving identities between integrals are
the change of variables and the Stokes formula.
Let us formalize them for the case of periods.
\begin{definition}
The space $P_+$ of effective periods is defined as a vector space over $\mathbb{Q}$
generated by the symbols $[(X,D,\omega,\gamma)]$ representing equivalence
classes of quadruples as above, modulo the following relations:
1) (linearity) $[(X,D,\omega,\gamma)]$ is linear in both $\omega$ and
$\gamma$
2) (change of variables) If $f:(X_1,D_1){\longrightarrow} (X_2,D_2)$
is a morphism of pairs defined over $\mathbb{Q}$,
$\gamma_1\in H_d(X_1(\mathbb{C}),D_1(\mathbb{C});\mathbb{Q})$ and $\omega_2\in \Omega^d(X_2)$ then
$$[(X_1,D_1,f^*\omega_2, \gamma_1)]=[(X_2,D_2,\omega_2,f_*(\gamma_1))]$$
3) (Stokes formula) Denote by $\tilde{D}$ the normalization of $D$
(i.e. locally it is the disjoint union of irreducible
components of $D$), the variety $\tilde{D}$ containing
a divisor with normal crossing $\tilde{D}_1$
coming from double points in $D$. If $\beta \in \Omega^{d-1}(X)$
and $\gamma \in H_d(X(\mathbb{C}),D(\mathbb{C});\mathbb{Q})$ then
$$[(X,D,d\beta,\gamma)]=[(\tilde{D},\tilde{D}_1,
\beta_{|\tilde{D}},\partial\gamma)]$$
where $\partial:H_d(X(\mathbb{C}),D(\mathbb{C});\mathbb{Q}){\longrightarrow} H_{d-1}(
\tilde{D}(\mathbb{C}),\tilde{D}_1(\mathbb{C});\mathbb{Q})$ is the boundary operator.
\end{definition}
It is conjectured in number theory that the evaluation homomorphism
$P_+{\longrightarrow} \mathbb{C}$ is a monomorphism, i.e. all identities between periods
can be proved using standard rules only. For example, the fact
that number $\pi$ is transcendental follows from this conjecture.
The effective periods form an algebra because the product of integrals
is again an integral (Fubini formula). The field of algebraic numbers
${\overline\mathbb{Q}}\subset \mathbb{C}$ can be considered as a subalgebra over $\mathbb{Q}$ of
the algebra $P_+$. An algebraic number $x\in {\overline \mathbb{Q}}$ which solves
a polynomial equation $P=0,\,\,\,P\in \mathbb{Q}[t]$, is the period of
$0$-dimensional variety $X\subset \mathbb{A}^1_{\mathbb{Q}}$ defined by the equation $P=0$.
The number $x$ gives a complex point of $X$, i.e. a $0$-chain.
The standard coordinate $t$ on the affine line $\mathbb{A}^1_{\mathbb{Q}}$ gives a
zero-form after restriction to $X$ whose pairing with
$x\subset X(\mathbb{C})\subset \mathbb{A}^1_{\mathbb{Q}}(\mathbb{C})=\mathbb{C}$ is tautologically equal to $x$.
It is convenient to extend the algebra of effective periods to a
larger algebra $P$ by inverting formally the element whose evaluation in
$\mathbb{C}$ is $2\pi i$. Informally, we can write that the whole algebra of periods
$P$ is $P_+[(2\pi i)^{-1}]$.
The algebra $P$ is an infinitely generated algebra over $\mathbb{Q}$, but as any
algebra it is an inductive limit of finitely-generated subalgebras.
This means that $\mathsf{ Spec}(P)$ is a projective limit of
finite-dimensional affine schemes over $\mathbb{Q}$. We claim that
$\mathsf{Spec}(P)$ carries a natural structure of a pro-algebraic
torsor over $\mathbb{Q}$.
The formula for the structure map $\Delta:P{\longrightarrow} P\otimes P\otimes P$
can easily be written in terms of period matrices of individual algebraic
varieties. Namely, let $(P_{ij})$ be the period matrix of an algebraic variety
consisting of pairings between classes running through
a basis in $H_*(X(\mathbb{C}),\mathbb{Q})$ and a basis in $H^*_{\mathrm{de \,Rham}}(X)$.
More generally, one should consider homology and cohomology of pairs
of algebraic varieties over $\mathbb{Q}$. It follows from several results in algebraic
geometry that the period matrix is a square matrix with entries
in $P_+$, and determinant in $\sqrt{\mathbb{Q}^{\times}}\cdot (2\pi i)^{\mathbb{Z}_{\ge 0}}$.
This implies that the inverse matrix has coefficients in the extended
algebra $P=P_+[(2 \pi i)^{-1}]$.
\begin{definition}
The triple coproduct in $P$ is defined by
\begin{equation}
\Delta (P_{ij}):=\sum_{k,l} P_{ik}\otimes (P^{-1})_{kl}\otimes P_{lj}
\end{equation}
for any period matrix $(P_{ij})$.
\end{definition}
We show how to calculate the triple coproduct in a simple example.
Let $X:=\mathbb{A}^1_{\mathbb{Q}}\setminus \{0\}$ be the affine line with deleted
point $0$, and $D:=\{1,2\}\subset X$ be a divisor in $X$.
The first homology group of pair (relative homology)
$$H_1(X(\mathbb{C}),D(\mathbb{C});\mathbb{Q})=H_1(\mathbb{C}\setminus\{0\},\{1,2\};\mathbb{Q})$$
is two-dimensional and is generated by two chains: a small closed
anti-clockwise oriented path around $0$, and the interval $[1,2]$.
The algebraic de Rham cohomology group $H^1_{\mathrm{de\,Rham}}(X,D)$
is also two-dimensional, and is generated by the $1$-forms $dt$ and $dt/t$
where $t$ is the standard coordinate on $X\subset \mathbb{A}_{\mathbb{Q}}^1$.
The period matrix is
\begin{equation}
\pmatrix{1 & \log(2)\cr 0 & 2 \pi i\cr}.
\end{equation}
{}From this one can deduce that
$$\Delta(2\pi i)=2\pi i \otimes {1\over 2\pi i}\otimes 2\pi i \,\,,$$
$$\Delta(\log(2))=\bigl(\log(2)\otimes {1\over 2\pi i}\otimes 2\pi i\bigr)
-\bigl(1\otimes {\log(2)\over 2\pi i}\otimes 2\pi i
\bigr)+ \bigl(1\otimes 1\otimes \log(2)\bigr )\,\,\,.$$
It is not clear why the definition given above is consistent, because
it is not obvious why the triple coproduct preserves the defining relations
in $P$. This follows more or less automatically from the following result:
\begin{theorem}[M.~Nori]
Algebra $P$ over $\mathbb{Q}$ is the algebra of functions on the pro-algebraic
torsor of isomorphisms between two
cohomology theories, the usual topological cohomology theory
$$H_{\mathrm{Betti}}^*:X\mapsto H^*(X(\mathbb{C}),\mathbb{Q})$$
and the algebraic de Rham cohomology theory
$$H_{\mathrm{de\,Rham}}^*:X\mapsto {\bf H} ^*(X,\Omega^*_X)$$
\end{theorem}
The motivic Galois group in Betti realization $G_{M,\mathrm{Betti}}$
is defined as the pro-algebraic group acting on $\mathsf{Spec}(P)$ from the
side of Betti cohomology. Analogously one defines the de Rham version
$G_{M,\mathrm{de\,Rham}}$. The category of motives is defined as the
category of representations of the motivic Galois group. It does not matter
which realization we choose because the categories for both
realizations can be canonically identified with each other.
Here is the ``symmetric'' definition of the category of motives:
\begin{definition}
The symmetric monoidal category of motives over $\mathbb{Q}$ with coefficients in $\mathbb{Q}$
is defined as the category of vector bundles on $\mathsf{Spec}(P)$ endowed
with two commuting actions of the motivic Galois groups
$G_{M,\mathrm{Betti}}$ and $G_{M,\mathrm{de\,Rham}}$.
\end{definition}
The following elementary definition gives a category equivalent
to the category of motives:
\begin{definition}
A framed motive of rank $r\ge 0$ is an invertible $(r\times r)$-matrix
$(P_{ij})_{1\le i,j,\le r}$ with coefficients in the algebra $P$,
satisfying the equation
\begin{equation}
\Delta (P_{ij})=\sum_{k,l} P_{ik}\otimes (P^{-1})_{kl}\otimes P_{lj}
\end{equation}
for any $i,j$. The space of morphisms from one framed motive to another,
corresponding to matrices
$$P^{(1)}\in GL(r_1,P),\,\,\,P^{(2)}\in GL(r_2, P),$$
is defined as
$$\{ T\in \mathsf{Mat}(r_2\times r_1,\mathbb{Q})| \,T P^{(1)}=P^{(2)}T\}$$
\end{definition}
The cohomology groups of varieties over $\mathbb{Q}$ can be considered as objects
of the category of motives. From comparison isomorphisms in algebraic
geometry follows that there are also $l$-adic realizations motives,
on which the absolute Galois group
$\mathsf{Gal}({\overline{\mathbb{Q}}}/\mathbb{Q})=\mathsf{Aut}({\overline{\mathbb{Q}}})$ acts.
The usual Galois group $\mathsf{Gal}({\overline{\mathbb{Q}}}/\mathbb{Q})$ is a
profinite group, projective limit of finite groups. It is the image of the
motivic Galois group $G_{M,\mathrm{Betti}}$ in motives coming from $0$-th
cohomology groups of schemes defined over $\mathbb{Q}$. It can be described also
as follows: the subalgebra ${\overline{\mathbb{Q}}}$ of $P$ is closed under
the triple coproduct,
$$\Delta_{|{\overline{\mathbb{Q}}}}
:{\overline{\mathbb{Q}}}{\longrightarrow} {\overline{\mathbb{Q}}}\otimes_{\mathbb{Q}}{\overline{\mathbb{Q}}}
\otimes_{\mathbb{Q}}{\overline{\mathbb{Q}}}$$
and its spectrum is a quotient torsor of $P$ on which
$\mathsf{Gal}({\overline{\mathbb{Q}}}/\mathbb{Q})$ acts simply transitively.
The definition of the category of motives given above
is a natural generalization of the ``folklore''
definition of the category of pure motives given
for example in [S]. There is an elaborated hypothetical picture
of motives in number theory, from which follows that
the category of motives as defined here is exactly what is expected.
The definition given here is equivalent
to the one advocated by M.~Nori in [N]. Specialists in motives consider
this definition as ``cheap'', and expect in the future
something more elaborate and not directly referring to
explicit realization functors (like Betti realization etc.).
The advantage of the definition given here is that it does not assume
the validity of any conjecture, and is directly applicable to
the present study of operads in deformation quantization.
\subsection{Grothendieck-Teichm\"uller group }
All the integrals appearing in situations listed in 4.1 share several
common features. The corresponding motives are so called
\textit{mixed Tate motives}, which means in particular that the
period matrix is upper-triangular in certain bases, with integral powers
of $(2\pi i)$ on the diagonal (like in the example in the previous
subsection). Also these motives are \textit{unramified over}
$\mathsf{Spec}(\mathbb{Z})$. This property can be expressed in terms of $l$-adic
representations of $\mathsf{Gal}({\overline{\mathbb{Q}}}/\mathbb{Q})$ and can be verified
in concrete situations by checking that certain discriminants are equal
to $0$,$1$ or $-1$. The period matrix (17) above,
is \textit{not} unramified. Namely, it is ramified at prime $2$.
We denote by $P_{\mathbb{Z},\mathrm{Tate}}$ the subalgebra of $P$ generated
by $(2\pi i)^{\pm 1}$ and by periods of mixed Tate motives unramified
over $\mathsf{Spec}(\mathbb{Z})$. There is a conjectural picture for it:
\begin{conjecture}
The quotient of the motivic Galois group (in the de Rham realization)
acting simply transitively on $\mathsf{Spec}(P_{\mathbb{Z},\mathrm{Tate}})$ is
a pro-solvable connected group over $\mathbb{Q}$, an extension of the multiplicative
groups scheme ${\bf G_m}=GL(1)$ by a pro-nilpotent
group whose Lie algebra is free and generated by elements
of weights $3,5,7,\dots$ (one element in each odd weight $\ge 3$)
with respect to the adjoint action of ${\bf G_m}$.
\end{conjecture}
This conjecture follows from general Beilinson conjectures
on motives and $K$-theory (see [Ne]).
There are two other conjectures concerning $P_{\mathbb{Z},\mathrm{Tate}}$.
\begin{conjecture}
$P_{\mathbb{Z},\mathrm{Tate}}$ is the subalgebra of $P$ generated by
$(2\pi i)^{\pm 1}$ and by periods whose evaluation in $\mathbb{C}$ are integrals
$I_{\epsilon_1,\dots, \epsilon_n}$ which appear in Drinfeld's associator
(see 4.1.A).
\end{conjecture}
There are two reasons for this conjecture. First of all, if we believe
in the picture of $P_{\mathbb{Z},\mathrm{Tate}}$ explained above,
there are $O(c^n)$ linearly independent effective unramified Tate motives
of weight $n$, where $c=1.32471...$ is the positive root of the equation
$x^3=x+1$. On the other hand, there are $2^{n-2}$ integrals
$I_{\epsilon_1,\dots, \epsilon_n}$, and with high probability
(because $c<2$) there are enough integrals to span the whole algebra
$P_{\mathbb{Z}, \mathrm{Tate}}$. The second reason is also probabilistic,
computer experiments confirm that the Poincar\'e series of the algebra
generated by integrals graded by weights
$w(I_{\epsilon_1,\dots, \epsilon_n}):=n$ equals
to the expected series $1/(1-t^2-t^3)$ up to $O(t^{13})$.
The next conjecture concerns the so called Grothendieck-Teichm\"uller
group $GT$ (see [D]).
\begin{conjecture}
The quotient of the motivic Galois group
acting simply transitively on the spectrum of the subalgebra
of $P$ generated by $(2\pi i)^{\pm 1} $ and integrals
$I_{\epsilon_1,\dots, \epsilon_n}$, coincides with the group $GT$.
\end{conjecture}
The group $GT$ is a pro-algebraic group over $\mathbb{Q}$, an extension of
${\bf G_m}$ by a pro-nilpotent group. One of the definitions of $GT$ is
as the group of automorphisms of the tower of pro-nilpotent completions
of pure braid groups $\pi_1(\mathbb{C}^n\setminus \mathrm{Diag})$ for all $n$.
The group of automorphisms of these pro-nilpotent completions
coincides with the group of automorphisms of the tower
of \textit{rational homotopy types} of classifying spaces of these groups.
Classifying spaces are $(\mathbb{C}^n\setminus \mathrm{Diag})_{n\ge 2}$,
configuration spaces of $\mathbb{R}^2=\mathbb{C}$. By general reasons the motivic Galois
group acts on the tower of rational homotopy types of
$(\mathbb{C}^n\setminus \mathrm{Diag})=
\left(\mathbb{A}_{\mathbb{Q}}^n\setminus \mathrm{Diag}\right)(\mathbb{C})$,
thus it maps to $GT$. Moreover, it is easy to see that the
periods which appear in the image of this action are exactly the
numbers $I_{\epsilon_1,\dots, \epsilon_n}$.
A priori, there is no reason for the tower of Malcev completed
braid groups to have a non-trivial automorphism.
Number theory provides (via motivic Galois group)
a supply of such automorphisms, and the conjecture 4 above says
that there is no other automorphism.
The group $GT$ maps to the group of automorphisms in homotopy sense
of the operad $\mathsf{Chains}(C_2)$. Moreover, it seems to coincide with
$\mathsf{Aut}(\mathsf{Chains}(C_2))$ when this operad is considered as an
operad not of complexes but of differential graded cocommutative coassociative
coalgebras (strictly speaking there is no coproduct on singular
cubical chains, but one can overcome this technical problem).
{}From now on we shall assume the validity of conjectures 2,3, and 4
and identify the group acting simply transitively on $P_{\mathbb{Z},\mathrm{Tate}}$
with the Grothendieck-Teichm\"uller group $GT$.
\subsection{Conjectures about torsors}
Remind that the integrals described in situations A)-G) in 4.1 give
complex points of the corresponding torsors, i.e. homomorphisms of algebras
over $\mathbb{Q}$
$$\mathcal{O}(T){\longrightarrow} \mathbb{C}\,\,\,.$$
Moreover, the values of the integrals in all these cases are periods,
and we proved that quadratic equations are satisfied using only the
Stokes formula. This implies that we have in fact a homomorphism
\begin{equation}
\mathcal{O}(T){\longrightarrow} P_{\mathbb{Z},\mathrm{Tate}}\,\,(\hookrightarrow P)
\end{equation}
In terms of pro-algebraic affine schemes, this means that we have a map
\begin{equation}
\mathsf{Spec}(P_{\mathbb{Z},\mathrm{Tate}})=
{\underline {\mathsf{Iso}}}_{\mathbb{Z},\mathrm{Tate}}
(H_{\mathrm{de\,Rham}}^*,H_{\mathrm{Betti}}^*){\longrightarrow} T
\end{equation}
between two torsors considered just as pro-algebraic schemes.
For example, in the formality theorem (cases D),E),F)) the torsor $T$
is the torsor of isomorphisms between two homotopy Lie algebras
$$T={\underline {\mathsf{Iso}}}(T^*_{\mathrm{poly}}, D^*_{\mathrm{poly}})$$
(see [K1] for a precise definition of $T^*_{\mathrm{poly}}$ and
$D^*_{\mathrm{poly}}$; one can safely replace $D^*_{\mathrm{poly}}$ by the
shifted by $[1]$ Hochschild complex of the algebra
$A:=\mathbb{R}[x_1,\dots,x_n]$, and $T^*_{\mathrm{poly}}$ by the algebra of
polynomial polyvector fields on $\mathbb{R}^d$). It is natural to ask whether
the constructed map is in fact a map of torsors.
\begin{conjecture}
In cases A),B),C),D),F) the map from $P$ to the corresponding torsor of
isomorphisms is a map of torsors.
\end{conjecture}
I checked that the generator of $\mathbb{Z}/2\mathbb{Z}\subset G_{M,\mathrm{Betti}}$
(acting via complex conjugation on $X(\mathbb{C})$ where $X$
is defined over $\mathbb{Q}$) corresponds to the natural involution
of the Hochschild complex $C^*(A,A)$ for $A=\mathbb{R}[x_1,\dots,x_n]$,
the Hochschild complex being considered as a dg-Lie algebra.
The involution acts on $\phi\in C^n(A,A)$ as
$$\phi\mapsto {\overline \phi},\,\,\,
{\overline \phi}(a_1\otimes\dots \otimes a_n):=
(-1)^{(n+1)(n+2)/2} \phi(a_n\otimes\dots \otimes a_1)\,\,\,.$$
In terms of deformation theory this corresponds to the passage
from a product $\star$ to the \textit{opposite} product
$$a{\overline \star} b:=b\star a$$
Also, the group ${\bf {G_m}}$ acts on $\mathsf{Spec}(P_{\mathbb{Z},\mathrm{Tate}})$
from the de Rham side, and also on
${\underline {\mathsf{Iso}}}(T^*_{\mathrm{poly}}, D^*_{\mathrm{poly}})$
from the side of $T^*_{\mathrm{poly}}$ rescaling polyvector fields
according to $\mathbb{Z}$-grading. Again, I checked that the two actions of
${\bf {G_m}}$ are compatible.
In cases the E),G) not listed in the conjecture, the values of the
integrals are \textit{real} numbers. It seems that the map between torsors
does not respect torsor structure. Moreover, I think that it is equal to
the composition of a map of torsors with a certain universal
map of schemes $P_{\mathbb{Z},\mathsf{Tate}}{\longrightarrow} P_{\mathbb{Z},\mathsf{Tate}}$
which is \textit{not} a map of torsors.
\subsection{Incarnations of the Grothendieck-Teichm\"uller group }
I present here two examples in which one can see the action of $GT$ on
deformation quantizations. Proofs will appear elsewhere.
\begin{theorem}
Let $\mathfrak{g}$ be a finite-dimensional Lie algebra over $\mathbb{R}$.
In the Duflo-Kirillov isomorphism (see [K1]) between the center
of the universal enveloping algebra $U\mathfrak{g}$, and the algebra of
invariant polynomials on $\mathfrak{g}^*$, one can replace the formal series
$$F(x)= \sqrt{e^{x/2}-e^{-x/2}\over x}$$
by the product
$$F_{\mathrm{new}}(x)=F(x)\cdot
\exp\left(\sum_{k=0}^{\infty} a_{2k+1} x^{2k+1}\right)$$
where $a_1,a_3,a_5,\dots$ are arbitrary constants.
For any such choice one gets again an isomorphism
compatible with products.
\end{theorem}
In particular, one can replace $F$ by the following entire function
$$F_{\mathrm{nice}}(x)={1\over{\Gamma}\bigl({x\over 2\pi i}+1\bigr)}\,\,\,$$
for the choice
$$a_1={{\rm Euler \,\,\,constant}\over 2 \pi i}
={0.57721...\over 2\pi i},\,\,a_3={\zeta(3)
\over 3(2\pi i)^3},\,a_5={\zeta(5)\over 5 (2\pi i)^5},\dots
$$
The Euler constant is (probably) not a period, and enters the
formula only for \ae sthetical reasons. The terms with coefficients
$a_3,a_5,\dots$ appear because of the action of $GT$.
As a corollary to Theorem 7 above we have the following
\begin{theorem}
If $\mathfrak{g}$ is a finite-dimensional Lie algebra,
then the differential operators on $\mathfrak{g}^*$ with constant coefficients
which are Fourier transforms of the following polynomials on $\mathfrak{g}$:
$$P_{2k+1}(\gamma):=\mathrm{Trace} (\mathrm{ad}(\gamma)^{2k+1}),\,\,\,k\ge 0$$
act on the subalgebra of $\mathrm{ad}(\mathfrak{g})$-invariant polynomials as
derivations, i.e. satisfy the Leibniz rule.
\end{theorem}
Unfortunately, as I learned from M.~Duflo, for all finite-dimensional
Lie algebras, operators such as in those considered in Theorem 7 above,
when restricted to $\mathsf{Sym}(\mathfrak{g})^{\mathfrak{g}}$ are all equal to zero.
Thus, we do not get anything visible here. Nevertheless, my result works
also for Lie algebras in rigid symmetric monoidal categories, for example
for a finite-dimensional Lie superalgebra. There is a chance that one
can find non-trivial examples there.
Another incarnation of the motivic Galois group is in a sense
Fourier dual to the previous one. Namely, let $X$ be a complex manifold
(or a smooth algebraic variety over a field of characteristic zero).
Define the Hochschild cohomology of $X$ as the following graded
commutative associative algebra:
$$HH^k(X):=\oplus_{i+j=k} H^i(X,\wedge^j(T_X))\,\,\,.$$
The product on $HH^*(X)$ is given by the usual cup-product
of polyvector fields and of cohomology classes.
Every element of the \textit{Hodge} cohomology
$$\oplus_{i,j} H^i(X,\wedge^j T^*_X)$$
gives a linear operator on $HH^*(X)$. It comes from the
cup-product in cohomology and from the convolution operator
$$\wedge^a T^*_X\otimes\wedge^b T_X{\longrightarrow} \wedge^{b-a} T_X$$
acting fiberwise on the level of bundles.
Operators corresponding to elements in $H^i(X,\wedge^i T^*_X)$
act on $HH^*(X)$ preserving the $\mathbb{Z}$-grading.
One can construct a version of characteristic classes
of vector bundles on $X$ with values in the diagonal part
$\oplus_{i\ge 0} H^i(X,\wedge^i T^*_X)$ of the Hodge cohomology.
This can be done using the Atiyah class of a vector bundle $\mathcal{E}$
taking values in $H^1(X,T^*_X\otimes \mathsf{End}(\mathcal{E}))$.
This class is the class of the extension of $\mathcal{E}$ by
$T^*_X\otimes {\mathcal{E}}$ given by the bundle of $1$-jets of sections
of $\mathcal{E}$. The traces of powers of the Atiyah class give
characteristic classes associated with monomial symmetric functions,
and other characteristic classes are polynomials in basic ones.
\begin{theorem}
Operators on $HH^*(X)$ corresponding to odd components
$$ch_{2k+1}(T_X)\in H^{2k+1}(X,\wedge^{2k+1}T^*_X)$$
of the Chern character $ch(T_X)$ of the tangent bundle,
are derivations of $HH^*(X)$ with respect to the cup-product.
\end{theorem}
The statement of this theorem is very easy for $k=0$, but
I do not know elementary proofs for higher values of $k$.
In the case of $X$ being a compact Calabi-Yau variety
one can identify the vector spaces $HH^*(X)$ and the usual
cohomology $H^*(X)$ using a Hodge decomposition and the convolution
with a volume element on $X$. The operator corresponding
to the class $\mathrm{ch}_1(T_X)=c_1(T_X)=0$ is zero, but higher operators
are in general non-zero. They correspond after the identification
of $HH^*(X)$ with $H^*(X)$ to the multiplication operators
by $\mathrm{ch}_{2k+1}(T_X)$.
In general, the result is that the algebra $HH^*(X)$ has
an infinite set of commuting automorphisms labeled by odd positive
integers, and it is a module over a solvable quotient of the
motivic Galois group. The Lie algebra of this quotient has the
following basis and bracket:
$$L_0,Z_3,Z_5,Z_7,\dots,\,\,\,[L_0,Z_{2k+1}]=(2k+1)Z_{2k+1},\,\,\,
[Z_{2k+1},Z_{2l+1}]=0\,\,\,.$$
\section{Relations to quantum field theories }
\subsection{Local fields and $d$-algebras }
There is no satisfactory definition yet of a quantum field theory (QFT).
One expects at least that a QFT on a manifold $X$
(``the space-time'') in the Euclidean framework
gives a super vector bundle of ``local fields'' $\Phi$ on $X$
(the bundle $\Phi$ may be infinite-dimensional), and correlators
$$\langle\phi_1(x_1)\dots \phi_n(x_n)\rangle \in \mathbb{C},\,\,\,n\ge 0$$
which are even polylinear maps from the tensor product of fibers of
$\Phi$ at pairwise distinct points
$$\langle\dots\rangle: \Phi_{x_1}\otimes\dots\otimes \Phi_{x_n}{\longrightarrow} \mathbb{C}$$
depending smoothly and $S_n$-equivariantly on
$$(x_1,\dots,x_n)\in \mathsf{Conf}_n(X)=X^n\setminus \mathrm{Diag}$$
Also one expects that there is an \textit{operator product expansion}
of fields at points converging to one point.
I am not going to discuss here in details the properties of the
operator products expansion.
It seems very plausible that with an appropriate definition
one can prove the following
\begin{conjecture}
For any QFT on vector space $V=\mathbb{R}^d$ with invariance under the action
of the group $G_d$ of parallel translations and dilatations
(in particular, for any conformal field theory), on the tensor product
$$\Phi_0\otimes \wedge^*(V^*)$$
there is a structure of a $d$-algebra.
\end{conjecture}
This implies that translation-invariant differential forms on $\mathbb{R}^d$
with values in local fields form
a homotopy Lie algebra $\mathfrak{g}^*$ whose graded components are
$$\mathfrak{g}^k=\Phi_0\otimes \wedge^{k+d-1}(V^*),\,\,\,k\in \{-(d-1),\dots,+1\}.$$
The moduli space associated with this homotopy
Lie algebra is unobstructed because $\mathfrak{g}^2=0$,
and it can be interpreted as the formal deformation theory of our QFT
as a translation-invariant theory,
with all renormalizations and regularizations automatically
included in the structure of $L_{\infty}$-algebra on $\mathfrak{g}^*$.
Presumably, the construction of (higher) brackets
on $\mathfrak{g}^*$ is closely related with the Hopf algebra
studied by Connes and Kreimer (see [CK]), where the
case of free massless theory is considered.
Also, it seems that the notion of an action of $(d+1)$-algebra
on a $d$-algebra is closely related with field theories
on manifolds with boundaries. This subject became very popular
in modern string theory after the Maldacena conjecture
on boundary conformal field theories for QFT on anti-de-Sitter
spaces (Lobachevsky spaces).
Here I should notice that a work of Mosh\'e and co-authors
(see [FFS] and references therein)
was one of the predecessors of the modern AdS picture.
\subsection{Action of the motivic Galois group on the moduli space of QFTs}
In the construction of perturbations
of a given conformal field theory one needs to calculate
(and often regularize) integrals of correlators over configuration spaces.
In the free theory case the integrals are exactly Feynman integrals
in the diagrammatic expansion.
D.~Broadhurst and D.~Kreimer (see [BK])
observed that all Feynman diagrams up to 7 loops
in any QFT in \textit{even dimensions}
gives same numbers as appear in Drinfeld associator. It is not clear
a priori why this
happens. In any case one can see immediately from formulas that
all constants are in fact periods.
\begin{conjecture} The motivic Galois group $G_{M,Betti}$
acts (in homotopy sense) on the homotopy
Lie algebra $\mathfrak{g}^*$ associated with the free massless theory
in any dimension. In the case of even dimension the action
factors through the quotient group $GT$ as in 3.4.
The action should be somehow related with the action
on values of Feynman integrals.
\end{conjecture}
I have only one ``confirmation'' of this conjecture,
that is the deformation quantization story.
In my proof of the formality theorem, the integrals which
appear in the explicit formula come from Feynman diagrams
for a perturbation of a free two-dimensional quantum field
theory on the Lobachevsky plane (see [CF]).
\bigskip
\noindent {\large\bf {Acknowledgements}}:
I would like to thank D.~Sternheimer and Y.~Soibelman for their
remarks and questions.
|
\section{Introduction}
The galaxy known as WLM is a low-luminosity, dwarf irregular galaxy in the
Local Group. A history of its discovery and early study was given by
Sandage \& Carlson (1985). Photographic surface photometry of the galaxy
was published by Ables \& Ables (1977). Its stellar population has been
investigated from ground-based observations by Ferraro et al (1989) and by
Minniti \& Zijlstra (1997). The former showed that the main body of the
galaxy consists of a young population, which dominates the light, while the
latter added the fact that there appears to be a very old population in
its faint outer regions. Cepheid variables were detected by Sandage \&
Carlson (1985), who derived its distance, and were reanalyzed by Feast \&
Walker
(1987) and by Lee et al. (1992). The latter paper used $I$ photometry of
the Cepheids and the RGB (red giant branch) distance criterion to conclude
that the distance modulus for WLM is 24.87 $\pm$ 0.08. The extinction
determined by Feast \& Walker (1987) is $A_B$ = 0.1.
Humason et al. (1956), when measuring the radial velocity of WLM, noticed a
bright
object next to it that had the appearance of a globular cluster. Its radial
velocity was the same as that of WLM, indicating membership. Ables \&
Ables (1977) found that the cluster's colors were like those of a globular
cluster, and Sandage \& Carlson (1985) confirmed this. Its total
luminosity is unusual for its being the sole globular of a galaxy. Sandage
\& Carlson (1985) quote a magnitude of $V$ = 16.06, indicating an absolute
magnitude of $M_V$ = -8.8.
This can be compared to the mean absolute magnitude of globulars in
galaxies, which is $M_V = -7.1 \pm 0.43$ (Harris 1991). The cluster, though
unusually bright, has only a small fraction of the $V$ luminosity of the
galaxy, which is 5.2 magnitudes brighter in $V$.
One could ask the question of whether there are other massive clusters in
the galaxy, such as luminous blue clusters similar to those in the
Magellanic Clouds. Minniti \& Zijlstra (1997), using the NTT and thus
having a wider field than ours, searched for other globular clusters and
found none. However, the central area of the galaxy has one very young,
luminous cluster, designated C3 in Hodge, Skelton and Ashizawa (1999). This
object is the nuclear cluster of one of the brightest HII regions (Hodge
\& Miller 1995). There do not appear to be any large intermediate age
clusters, such as those in the Magellanic Clouds or that recently
identified spectroscopically in the irregular galaxy NGC 6822 by Cohen \&
Blakeslee (1998) .
No other Local Group irregular galaxy fainter than $M_V$ = -16 contains a
globular cluster. The elliptical dwarf galaxies NGC 147 and NGC 185 (0.8
and 1.3 absolute magnitudes brighter than WLM, respectively) do have a few
globular clusters each and Fornax (1.7 absolute magnitudes fainter) has
five, which makes it quite anomalous, even for an elliptical galaxy (see
Harris, 1991, for references).
Another comparison can be made using the specific frequency parameter, as
defined and discussed by Harris (1991). The value of the specific frequency
calculated for WLM is 7.4, which can be compared to Harris' value
calculated for late-type galaxies, which is $0.5 \pm 0.2$. The highest
average specific frequency is found for nucleated dwarf elliptical galaxies
by Miller et al (1998), which is $6.5 \pm 1.2$, while non-nucleated
dwarf elliptical galaxies have an average of $3.1 \pm 0.5$. These values
are similar to those found by Durrell et al (1996), implying that the
specific frequency for WLM is comparable to that for dwarf elliptical
galaxies but possibly higher than that for other late-type galaxies.
Because the WLM cluster as a globular in an irregular dwarf galaxy is
unique, it may represent an unusual opportunity to
investigate the question of whether Local Group dwarf irregulars share the
early history of our Galaxy and other more luminous Group members, which formed
their massive clusters some 15 Gyr ago, or whether they formed later, as
the early ideas about the globular clusters of the Magellanic Clouds seemed
to indicate. Of course, we now know that the LMC has several true globular
clusters that are essentially identical in age to the old halo globulars of
our Galaxy (Olsen et al. 1998, Johnson et al. 1998), so the evidence
suggesting a delayed formation now seems to come only from the SMC. In any
case, WLM gives us a rare opportunity to find the oldest cluster (and
probably the oldest stars) in a more distant and intrinsically much less
luminous star-forming galaxy in the Local Group.
\section{Data and Reduction}
\subsection{Observations}
As part of a Cycle 6 HST GO program, we obtained four images of the WLM
globular cluster on 26 September, 1998. There were two exposures taken
with the F814W filter of 2700 seconds and two with the F555W filter, one
each of 2700 seconds and 2600 seconds. The globular cluster was centered on
the PC chip and the orientation of the camera was such that the WF chips
lay approximately along the galaxy's minor axis, providing a representative
sample of the WLM field stars to allow us to separate cluster stars
reliably.
\subsection{Reductions}
With two images of equal time per filter, cosmic rays were cleaned with
an algorithm nearly identical to that used by the IRAF task CRREJ. The
two images were compared at each pixel, with the higher value thrown out
if it exceeded 2.5 sigma of the average. The cleaned, combined F555W
image is shown in Figure \ref{fig_image}.
\placefigure{fig_image}
Photometry was then carried out using a program specifically designed to
reduce undersampled WFPC2 data. The first step was to build a library
of synthetic point spread functions (PSFs), for which Tiny Tim 4.0
(Krist 1995) was used.
PSFs were calculated at 49 positions on each chip in F555W and F814W,
subsampled at 20 per pixel in the WF chips and 10 in the PC chip. The
subsampled PSFs were adjusted for charge diffusion and estimated subpixel
QE variations, and combined for various locations of a star's center
within a pixel. For example, the library would contain a PSF for the
case of a star centered in the middle of a pixel, as well as for a star
centered on the edge of a pixel. In all, a 10x10 grid of possible centerings
was made for the WF chips and a 5x5 grid for the PC chip.
This served as the PSF library for the photometry.
The photometry was run with an iterative fit that located stars and
then found the best combinations of stellar profiles to match the
image. Rather than using a centroid to determine which PSF to use for
a star, a fit was attempted with each PSF centered near the star's position
and the best-fitting PSF was chosen. This method helped
avoid the problem of centering on an undersampled image. Residual
cosmic rays and other non-stellar images were removed through a
chi-squared cut over the final photometry list.
The PSF fit was normalized to give the total number of counts from the
star falling within a 0.5 arcsec radius of the center. This count rate was
then converted into magnitudes as described in Holtzman et al (1995), using
the CTE correction, geometric corrections, and transformation.
For the color-magnitude diagram (CMD) and luminosity
function analyses below, roughly the central 20\% of the image was analyzed
to maximize the signal from the globular while minimizing the background
star contamination. In that region, the effect of background stars is
negligible.
The CMD from this method, showing all stars observed, is shown in Figure
\ref{fig_phot}a, with the same data reduced with DAOPHOT shown in Figure
\ref{fig_phot}b. Error bars are shown corresponding to the artificial star
results (which account for crowding in addition to photon statistics),
rather than the standard errors from the PSF fits and
transformations. The photometry list is given in Table \ref{tabphot},
which will appear in its entirety only in the electronic edition. The
table contains X and Y positions of each star, with $V$ and $I$ magnitudes
and uncertainties. Uncertainties given are from the PSF fitting and
from the transformations.
\placetable{tabphot}
\placefigure{fig_phot}
Artificial star tests were made, with each star
added and analyzed one at a time to minimize additional crowding often
caused by the addition of the artificial stars. The artificial
stars were added to both the combined $V$ and $I$ images, so that a library
of artificial star results for a given position, $V$ magnitude, and color
could be built. In addition to completeness corrections, these data were
employed in the generation of synthetic CMDs for the determination
of the star formation history of the cluster.
\section{Analysis}
\subsection{Luminosity Function}
The luminosity functions (LFs) of the globular cluster in $V$ and $I$ are
shown in Figure \ref{fig_lf}a and \ref{fig_lf}b, respectively, binned
into 0.5 magnitude bins. Theoretical luminosity functions are given
as well, from interpolated Padova isochrones (Girardi et al 1996, Fagotto et
al 1994) using the star formation parameters and distance obtained
in the CMD analysis below.
\placefigure{fig_lf}
The observed and theoretical LFs are in excellent agreement. The bump in
the observed $V$ LF between magnitudes 25 and 26, and the bump in the
observed $I$ LF starting at magnitude 24.5 are due to the horizontal
branch stars, which cannot be separated from the rest of the CMD
cleanly. The only other significant deviation, the bump in the $V$ LF
between magnitudes 22.5 and 23, is the clump of stars at the
tip of the RGB, which is also observed in the CMD. This seems to be a
statistical
fluke, a result of the relatively small number of stars in that part
of the RGB, and is similar to statistical flukes seen in Monte Carlo
simulations. Thus as far as can be determined, the observed LF agrees
with the theoretical expectations.
\subsection{Color-Magnitude Diagram}
For the CMD analysis, a cleaner CMD was achieved by omitting all stars
with PSF fits worse than a chi-squared value of 3.
The observed $V, V-I$ CMD is shown in Figure \ref{fig_cmd}a, and was
analyzed as described in Dolphin (1997). Interpolated Padova isochrones
(Girardi et al 1996, Fagotto et al 1994) were used to generate the
synthetic CMDs, with photometric errors and incompleteness simulated
by application of artificial star results to the isochrones.
No assumptions were made regarding the star formation history, metallicity,
distance, or extinction to the cluster, and a fit was attempted with all of
these parameters free. The best fits that returned single-population
star formation histories were then combined with a weighted average to
determine the best parameters of star formation. Uncertainties were
derived by taking a standard deviation of the parameters from the fits,
and thus include the fitting errors and uncertainties resulting from
an age-metallicity-distance ``degeneracy.'' Systematic errors due to the
particular choice of evolutionary models are naturally present, but are not
accounted for in the uncertainties. The following parameters were
obtained:
\begin{itemize}
\item Age: 14.8 $\pm$ 0.6 Gyr
\item Fe/H: -1.52 $\pm$ 0.08
\item Distance modulus: 24.73 $\pm$ 0.07
\item Av: 0.07 $\pm$ 0.06
\end{itemize}
A synthetic CMD constructed from these parameters is shown in Figure
\ref{fig_cmd}b, using the artificial star data to mimic photometric
errors, completeness, and blending in the data. The poorly reproduced
horizontal branch is a result of the isochrones we used, but the giant
branch was reproduced well, with the proper shape and position.
\placefigure{fig_cmd}
\subsection{Structure}
Profiles were calculated in bins of 10 pixels (0.45 arcsec) in both the
$V$ and $I$ images, and are shown in Figure \ref{fig_prof}, corrected for
incompleteness (both as a function of magnitude and position). The cutoff
magnitudes of 27 in $V$ and 26 in $I$ were chosen to minimize the corrections
required due to incompleteness.
Additionally, the central bin (0-10 pixels)
was omitted because of extreme crowding problems. The remaining
bins were fit to King models with a least-squares fit. The best
parameters for the King models (assuming a
distance modulus of 24.73) are as follows (shown by the solid lines in Figure
\ref{fig_prof}).
\begin{itemize}
\item core radius: 1.09 $\pm$ 0.14 arcsec (4.6 $\pm$ 0.6 pc)
\item tidal radius: 31 $\pm$ 15 arcsec (130 $\pm$ 60 pc)
\item core density: 59 $\pm$ 8 stars/arcsec$^2$ (3.2 $\pm$ 0.4 stars/pc$^2$) $V$, 44 $\pm$ 6 stars/arcsec$^2$ (2.4 $\pm$ 0.3 stars/pc$^2$) $I$
\item background density: 0.77 $\pm$ 0.12 stars/arcsec$^2$ (0.042 $\pm$ 0.007 stars/pc$^2$) $V$, 0.77 $\pm$ 0.12 stars/arcsec$^2$ (0.042 $\pm$ 0.007 stars/pc$^2$) $I$
\end{itemize}
For a distance modulus of 24.87 (Lee et al. 1992), the corresponding sizes
would be 7\% larger.
For comparison, Trager et al. (1993) find that 2/3 of Milky
Way clusters have core radii between approximately 5 and 60 pc.
\placefigure{fig_prof}
\section{Conclusions}
Our analysis shows that the WLM globular cluster is virtually
indistinguishable from a halo globular in our Galaxy. We find that a
formal fit to theoretical isochrones indicates an age of 14.8 $\pm$ 0.6 Gyr,
which agrees with ages currently being measured for Galactic globulars
(e.g., vandenBerg 1998) and a metallicity of [Fe/H] of -1.52 $\pm$ 0.08, a
typical globular cluster value that is similar to that obtained for the
outer field giant stars along the minor axis of WLM by Minniti and Zijlstra
(1997) and by us (Dolphin 1999). The distance modulus for the cluster,
derived independently from the parent galaxy, is 24.73 $\pm$ 0.07, which
agrees within the errors with that derived from Cepheids and the RGB (Lee
et al. 1992) of the galaxy.
In structure the globular is elongated in outline, with a mean radial profile
that fits a King (1962) model within the observational uncertainties. We
derive a core radius of 1.09 $\pm$ 0.14 arcsec and a tidal radius of 31 $\pm$
15 arcsec, which translate to 4.6 $\pm$ 0.6 pc and 130 $\pm$ 60 pc,
respectively. The core radius is very similar to that found for massive
globulars in our Galaxy (Trager et al. 1993), while the tidal radius,
though quite uncertain, is rather large by comparison. The former result
indicates that formation conditions in this galaxy near its conception were
such that a massive, highly concentrated star cluster could form, despite
the very small amount of the total mass of material available. The latter
result is probably an indication that the tidal force of the galaxy on the
cluster is small.
The presence of a normal, massive globular cluster in this dwarf irregular
galaxy may be an useful piece of evidence regarding the early history of
star, cluster and galaxy formation. Recent progress in the field of
globular cluster formation has resulted from both observational and
theoretical studies (Searle \& Zinn 1978, Harris \& Pudritz 1994,
McLaughlin \& Pudritz 1994, Durrell et al 1996, Miller et al 1998, and
McLaughin 1999). Although the uncertainties from a single data point are
sufficiently large to discourage quantitative analysis, the presence of
a globular cluster in WLM would constrain formation models that predict
$\ll 1$ cluster in such a galaxy.
\acknowledgements
We are indebted to the excellent staff of the Space Telescope Science
Institute for obtaining these data and to NASA for support of the analysis
through grant GO-06813.
\begin{table}
\dummytable\label{tabphot}
\end{table}
\clearpage
|
\section{Introduction}
Two main motivations for the low energy supersymmetry are the
alleviation of the hierarchy problem and the successful unification
of gauge couplings in the context of the minimal supersymmetric
standard model (MSSM) \cite{mssm}. The quest for supersymmetry
has become one of the central issues of present day particle
physics.
Today, on the other hand, we have growing observational evidence
in favor of non-vanishing neutrino masses. The most natural and
the simplest way to generate small neutrino masses is provided by
the see-saw mechanism \cite{seesaw}. The accumulation of solar
and atmospheric neutrino data points to the see-saw (right-handed
neutrino) mass scale $M_R$ in the range of $10^{10}-10^{15}$ GeV
\cite{smirnov99}. For believers in supersymmetry there is thus an
important question as to whether there can be an
intermediate scale
below the grand unification scale $M_X$. The
conventional wisdom in
the context of minimal grand unified theories
without extra fine-tuning seems to be no.
Recently, we have pointed out \cite{amrs99} an important connection
between the renormalizable see-saw mechanism and low energy
supersymmetry.
First, if B$-$L symmetry is spontaneously broken, we can show that
$R$-parity is an exact symmetry and the low-energy effective theory
is the MSSM. Second, in many cases supersymmetry is consistent
with the canonical form of the see-saw, an important achievement.
It is then essential to know whether we can have $M_R$ in the
intermediate regime, as may be needed by neutrino data. In this
letter we wish to show that, contrary to the conventional
wisdom, it may be quite natural
to have $M_R$ as an intermediate scale. In the minimal supersymmetric
SO(10) theory $M_R$ can be as low as $10^{13}$ GeV with the prediction
of lowering $M_X$ close to $10^{15}$ GeV, giving thus the old
$d=6$ proton decay mode $p\to\pi^0e^+$ as the possibly dominant one.
The reason for a different prediction from the minimal supersymmetric
SU(5) stems from the violation of the survival principle \cite{survival}
and it has generic features beyond any particular model. In other words,
supersymmetric GUTs beyond
SU(5) contain many potentially light supermultiplets which may
modify the usual gauge coupling running. This is an important,
albeit simple point, often overlooked in the literature (but not
always, see for example \cite{light,cceel97}). Since it is generic to
supersymmetric gauge theories, we address it more carefully and illustrate
it in a number of examples, starting from a minimal U(1) model, all the
way to the supersymmetric SO(10). In all cases we will find
a violation of the naive survival principle.
\section{Survival of the fittest}
The survival principle is the natural assumption that a particle in
multi scale theories takes the largest possible mass consistent with
the symmetries of the theory. In particular, this means that the
singlets under a particular gauge symmetry have masses that
correspond to the scale of the larger symmetry under which they are
not singlets. Of course, if they do not participate in symmetry breaking
and have a gauge invariant mass, then they can go to an
even higher scale.
An important aspect of this principle is the usual manifestation of the
Higgs mechanism in ordinary gauge theories. When a multiplet $\Phi$
develops a nonvanishing vacuum expectation value (VEV)
through the potential
\begin{equation}
V=-m^2\Phi^\dagger\Phi+\lambda^2(\Phi^\dagger\Phi)^2+...\;,
\label{vphi}
\end{equation}
\noindent
the Higgs scalars get masses of the order
\begin{equation}
m_\Phi\approx\lambda\langle \Phi\rangle
\end{equation}
\noindent
except for the unphysical Higgs and the possible pseudo-Goldstone
bosons in the case of accidental symmetries of the potential. Of
course, in general we have more self couplings in (\ref{vphi});
$\lambda$ stands generically for all of them. Notice that there is no
symmetry that forbids the coupling $\lambda$, in as much as there is no
symmetry that forbids the mass term.
This is intimately related to the problem of hierarchies and thus we
should not be surprised if the situation gets drastically different in
supersymmetric theories.
In other words, unless one resorts to the unnatural fine-tuning
$\lambda\ll 1$, we expect $m_\Phi\sim \langle \Phi\rangle$. This
looks rather natural, so natural, that it is normally taken for
granted even in supersymmetry. It is however potentially wrong as
we show now. In supersymmetry the $\Phi^4$ term in (\ref{vphi})
arises from the superpotential $W=\lambda\Phi^3/3+...$. However,
the cubic term is often forbidden for symmetry reasons and so
is the quartic self-coupling in the potential. In general there
will be higher dimensional terms of the form
\begin{equation}
W_{nr}={\Phi^{n+3}\over M^n}\;,
\end{equation}
\noindent
where $M$ is the cutoff scale of the theory; $n>0$ and it depends on
the symmetry in question. Thus the $\lambda$ coupling is strongly
suppressed, $\lambda_{\rm eff}\sim \langle \Phi\rangle^n/M^n$, and
\begin{equation}
m_\Phi\approx \lambda_{\rm eff}\langle \Phi\rangle
\approx{\langle\Phi\rangle^{n+1}\over M^n}\;.
\end{equation}
Even for $n=1$, we can have $m_\Phi\sim \langle
\Phi\rangle^2/M\ll \langle \Phi \rangle$,
since $\langle \Phi\rangle\ll M$. Here $\Phi$ stands for the whole
super-multiplet of bosons and fermions, since we will deal with
$\langle \Phi\rangle \gg M_W$ when supersymmetry is assumed to be a good
symmetry. This fact is clearly important for phenomenology, but
also for the unification of couplings. If $M\sim M_X$, and
$\langle \Phi\rangle=M_I$, the intermediate scale, many particles
will start contributing to running at much lower energies
$M_I^2/M_X$. In what
follows we give some examples of the above with growing
complexity and physical relevance.
\subsection{A U(1) prototype example}
Take the simplest supersymmetric anomaly-free U(1) gauge
symmetry with two chiral superfields $\Phi$ and
$\overline\Phi$ (charges $\pm 1$). The U(1) symmetry forbids cubic
couplings and the renormalizable superpotential is trivial,
$W = m\Phi\overline\Phi$. It implies no breaking of the symmetry,
$\langle \Phi\rangle = \langle\overline\Phi\rangle = 0$.
In order to have symmetry breaking, one needs at least a $d=4$ term
in the superpotential,
\begin{equation}
W = m\Phi\overline\Phi+{(\Phi\overline\Phi)^2\over 2M}\;,
\end{equation}
\noindent
which together with
\begin{equation}
V_D = {g^2\over 2}(|\Phi|^2-|\overline\Phi|^2)^2
\end{equation}
\noindent
provides nonvanishing VEVs: $\langle \Phi\rangle=\langle
\overline\Phi\rangle=\sqrt{mM}$.
Obviously, the effective coupling is suppressed
\begin{equation}
\lambda\sim \langle \Phi\rangle/M\;.
\end{equation}
Whereas the combination $\Phi-\overline\Phi$ belongs to the super-Higgs
multiplet (mass of order $g\langle \Phi\rangle$), $\Phi+\overline\Phi$
has a smaller mass
\begin{equation}
m_{(\Phi+\bar\Phi)} \approx \langle \Phi\rangle^2/M\;.
\end{equation}
This simple model portrays well the general
situation we will find in the following.
\subsection{The supersymmetric Left-Right model}
The minimal such model mimics completely the above U(1) example.
The gauge group is SU(2)$_L\times$ SU(2)$_R\times$ U(1)$_{B-L}$
\cite{leftright}. We have left handed ($\Delta$, $\overline\Delta$)
and right handed ($\Delta_c$, $\overline\Delta_c$) triplets and in
the absence of nonrenormalizable terms there is no interaction
in the superpotential \cite{amrs98}. Again, one needs nonrenormalizable
terms to generate symmetry breaking. Just as above, except for
the super-Higgs multiplet, the rest of the particles should have
masses of the order $M_R^2/M$, where
$M_R=\langle \Delta_c\rangle=\langle \overline\Delta_c\rangle$
is the symmetry breaking scale of parity and SU(2)$_R$ and $M$
is the cutoff.
The neutral multiplets in $\Delta_c$ and $\overline\Delta_c$
get a VEV, and the corresponding masses through the super-Higgs mechanism.
The states that do not belong to the super-Higgs multiplet are one
neutral complex combination and two doubly charged chiral multiplets
from $\Delta_c $ and $\overline\Delta_c$, and of course all the states
from $\Delta$ and $\overline\Delta$.
This leads
to an important prediction of light doubly charged supermultiplets
potentially observable by experiment \cite{ams97}, with striking
signatures at future colliders\cite{dm98}. This result
can be extended to include an arbitrary number of gauge singlet
fields \cite{cm97}. In this case the lightness of the doubly charged
multiplets can be traced to their pseudo-Goldstone nature \cite{cm97}.
Namely, with parity odd singlet(s) one can generate the VEVs for the
right handed fields, but the renormalizable superpotential has a larger
accidental symmetry and the doubly charged fields are the Goldstone
modes. They get masses if one includes the higher dimensional
terms which break the accidental symmetry.
However, our point is much more general and is valid whenever one
needs to invoke higher dimensional operators to generate relevant
interactions. It says simply, as we already emphasized, that
except for the super-Higgs mechanism, the rest of the supermultiplet
masses are suppressed by the cut-off scale associated with
the higher dimensional interactions. In the minimal model
(without singlets)
the lightness of the doubly charged states has nothing to
do with the pseudo-Goldstone mechanism: simply, in the absence
of higher dimensional operators there is no symmetry breaking
whatsoever. When they are included and L-R symmetry is broken,
the doubly charged states are light because they do not
belong to the super-Higgs multiplet.
\section{Supersymmetric SO(10) theory}
From the point of view of unification this will be the simplest
illustration of the above idea. The minimal renormalizable theory
requires the symmetric traceless ${\bf 54}$ supermultiplet
and the antisymmetric ${\bf 45}$ adjoint. This allows
for the breaking of SO(10) down to SU(2)$_L\times $SU(2)$_R\times$
U(1)$_{B-L}\times$ SU(3)$_C$. Further breaking is achieved by either
({\it i}) $ {\bf 126} $ and $ {\bf \overline{126}}$ or
({\it ii}) {\bf 16 } and ${\bf \overline{16}}$.
The former turns out to be more interesting so we study
it first in detail.
\subsection{${\bf 126}$ and {${\bf \overline{126}}$} or
renormalizable see-saw mechanism}
A renormalizable SO(10) theory with a see-saw requires the minimum
set of Higgs representations which break SO(10) down to the MSSM
\cite{leem94}
\begin{equation}
S = {\bf 54} \; , \quad A = {\bf 45} \; , \quad \Sigma = {\bf 126} \; ,
\quad \overline \Sigma = \overline{{\bf 126}} \;.
\label{higgses}
\end{equation}
Although SO(10) is anomaly-free, as is well-known, one has to use
both $\Sigma$ and $\overline \Sigma$ in order to ensure the flatness
of the D-piece of the potential at large scales $\gg M_W$. The most
general superpotential containing $S, A, \Sigma $ and $\overline \Sigma$
is
\begin{eqnarray}
W &= &{m_S \over 2} {\rm Tr}\, S^2 + {\lambda_S \over 3} {\rm Tr}\, S^3
+ {m_A \over 2} {\rm Tr}\, A^2 + \lambda {\rm Tr}\, A^2S \nonumber \\
& & + m_\Sigma \Sigma \overline\Sigma + \eta_S \Sigma^2 S +
\overline\eta_S
{\overline\Sigma}^2 S + \eta_A \Sigma\overline\Sigma A\;.
\label{superpot}
\end{eqnarray}
We wish to obtain the pattern of symmetry breaking
\begin{eqnarray}
SO(10) & \stackrel{\langle S \rangle = M_X }{\longrightarrow} &
SU(2)_L\times SU(2)_R\times SU(4)_C \nonumber \\
& \stackrel{\langle A \rangle = M_C }{\longrightarrow} &
SU(2)_L\times SU(2)_R\times U(1)_{B-L}\times SU(3)_C \nonumber \\
& \stackrel{\langle \Sigma \rangle = \langle {\overline
\Sigma} \rangle = M_R }{\longrightarrow} &
SU(2)_L\times U(1)_{Y}\times SU(3)_C
\label{sixteen}
\end{eqnarray}
\noindent
and this is achieved with non-vanishing VEVs in the directions
\begin{equation}
s = \langle (1,1,1)_S \rangle \quad
a = \langle (1,1,15)_A \rangle \quad
b = \langle (1,3,1)_A \rangle \quad
\sigma = \langle (1,3,10)_\Sigma \rangle \quad
\overline\sigma = \langle (1,3,\overline{10})_{\overline\Sigma}
\rangle \quad
\label{vevnames}
\end{equation}
\noindent
(where here and in the following we label fields by their
SU(2)$_L \times$ SU(2)$_R \times$ SU(4)$_C$ quantum numbers).
Notice that $\langle S \rangle$ also preserves the discrete
parity symmetry $D_{LR}$ \cite{parida}, which is then broken in the
next step by the parity-odd singlet in $(1,1,15)_A$. This will
of course cause the left-right inequality of masses already at
the scale $M_C$.
The F-flatness equations are
\begin{eqnarray}
m_S s + {1 \over 2} \lambda_S s^2 + {2\over 5}
\lambda (a^2 -b^2) &=& 0 \;, \nonumber \\
a (m_A + 2 \lambda s) + {1 \over 2} \eta_A \sigma\overline\sigma
= b (m_A - 3 \lambda s) + {1\over 2} \eta_A \sigma \overline\sigma
&=& 0 \;, \nonumber \\
\sigma \left[ m_\Sigma + \eta_A (3 a + 2 b) \right]
&=& 0
\label{efeqs}
\end{eqnarray}
\noindent
and $\sigma = \overline\sigma $ guarantees D-flatness.
In order to get the required multi-scale symmetry breaking one needs
$s\gg a \gg \sigma= \overline\sigma \gg b \simeq \sigma^2/s$.
This can be achieved by paying the usual price of fine-tuning
\begin{equation}
m_A + 2 \lambda s \simeq {\sigma^2 \over a} \ll s\;,
\label{finea}
\end{equation}
\noindent
which then ensures
\begin{equation}
b \simeq {\sigma^2 \over s} \ll \sigma
\label{finea2}
\end{equation}
\noindent
(it is important to keep in mind that $b$ can never vanish).
A comment is in order. In the F flatness conditions (\ref{efeqs})
we ignore the fields in {\bf 16} and {\bf 10}
dimensional representations.
This is justifiable for the Standard Model non-singlet fields,
but not for $\tilde{\nu^c}$ in {\bf 16}. However, in the vacuum
$\sigma = \overline \sigma$, $\langle \tilde{\nu^c}\rangle = 0$,
as can be easily proved following the similar analysis in the
context of LR theories \cite{amrs98}.
We now compute the spectrum of the
theory. This has been discussed
in reference\cite{leem94}, however only in the limit of a
single step breaking of SO(10). In this limit one misses the
possibility of light states which have masses suppressed by the
ratios of different scales of symmetry breaking. In what follows
we identify all such states together with the rest of the
spectrum\footnote{For previous attempts in supersymmetric
SO(10) theories with 126 fields see \cite{aulamoha,sato},
where in \cite{aulamoha} the presence of light states
was noticed. However, the effects of $M_I^2/M$ that are central
to our paper were missed.}.
At $M_X$, $(2,2,6)_S$ gets a mass of the order $\langle S\rangle$
due to the super-Higgs mechanism. So do all the remaining fields
in $S$ and almost all of $A$, through the terms Tr$S^3$ and Tr$SA^2$.
The sole exception
are the $(1,1,15)$ fields in $A$, which we choose to have a much smaller
mass imposing the fine-tuning condition (\ref{finea}): this multiplet
will thus get a VEV in the next stage of symmetry breaking, at a much
lower scale. Couplings with $S$ also give a
mass $\sim M_X$ to all fields in $\Sigma$ and $\overline \Sigma$,
with the exception of the SU(4) decuplets. There is no
fine-tuning involved here; this is automatic, since $S$ is
coupled only to $\Sigma^2$ or $\overline{\Sigma}^2$ and not to
$\Sigma\overline{\Sigma}$.
In the next stage $(1,1,15)_A$ will get a nonvanishing VEV in the
direction of the SU(3) color singlet. Again, the super-Higgs
mechanism gives a mass $M_C$ to the colored triplets in $(1,1,15)_A$.
The couplings with $A$ provide masses at this scale to the left-handed
components $(3,1,\overline{10})_\Sigma$ and $(3,1,10)_{\overline\Sigma}$,
and to the colored fields in their right-handed analogues. As mentioned
before the reason for the asymmetry in the left-right masses already
at this scale comes from the fact that the singlet in $(1,1,15)_A$
is parity-odd. But as we
anticipated, the color octet and the singlet in $A$ will survive the
Pati-Salam breaking. The reason once again can be traced to the
superpotential (\ref{superpot}). Both $A$ and the $\Sigma$
($\overline\Sigma $) fields lack cubic self-interactions.
In order to get effective interactions one has to integrate out the
heavy field $S$. Notice that the color octets may get a
contribution of the order $M_R^2/M_C$, which is just due to
the necessary fine-tuning of formula (\ref{finea}).
At this point the situation resembles the Left-Right model discussed
above. At the scale $M_R$, the SU(2)$_R$ triplets in $\Sigma$ and
$\overline\Sigma$ get their VEVs, but two doubly-charged chiral
multiplet and one neutral combination remain light. The mass spectrum
is summarized in the Table below.
\begin{center}
\framebox{\begin{tabular}{l|c}
\hspace{2cm}State & Mass \\
\hline
\begin{tabular}{l}
all of $S$ in ${\bf 54}$ \\
all of $A$ in ${\bf 45}$, except $(1,1,15)_A$\\
all of $\Sigma$ in {\bf 126} + $\overline \Sigma$
in ${\bf \overline{126}}$, except SU(4)$_C$ decuplets
\end{tabular}
& $\sim M_X$ \\
\hline
\begin{tabular}{l}
$(3,1,\overline{10})_\Sigma$ + $(3,1,{10})_{\overline\Sigma}$ \\
color triplets and sextets of $(1,3,{10})_\Sigma$ and
$(1,3,\overline{10})_{\overline\Sigma}$ \\
color triplets of $(1,1,15)_A$
\end{tabular}
& $\sim M_C$ \\
\hline
\begin{tabular}{l}
$(\delta_c^0 - \overline{\delta_c^0}), \delta_c^+,\overline{\delta_c^+} $\\
\hspace{0.5cm} from the color singlets of $(1,3,{10})_\Sigma$
and $(1,3,\overline{10})_{\overline\Sigma}$
\end{tabular}
& $\sim M_{R}$ \\
\hline
color octet and singlet of $(1,1,15)_A$
& $\sim Max\left[ {M_R^2 /M_C}, {M_C^2 /M_X} \right]$ \\
\hline
\begin{tabular}{l}
$(\delta_c^0 + \overline{\delta_c^0}) \; , \;\delta_c^{++},\overline{
\delta_c^{++}} $\\
\hspace{0.5cm} from the color singlets of $(1,3,{10})_\Sigma$
and $(1,3,\overline{10})_{\overline\Sigma}$
\end{tabular}
& $\sim {M_{R}^2 / M_X}$
\end{tabular}}
\vspace{0.5cm}
\end{center}
\vspace{0.5cm}
In addition, we have to include a {\bf 10}
representation, which contains the MSSM Higgs doublets. Couplings with $S$ (and the usual
fine-tuning) will ensure that the colored fields in {\bf 10} get
a mass $\sim M_X$.
However, as is well known, at least
two ten-dimensional Higgs representations are needed in general
in order to generate non-vanishing quark mixing angles in the
CKM matrix. This implies two extra Higgs doublets, which get
a mass through the VEV $b$ in $A$ of the order $M_R^2/M_X$.
The remaining fields in {\bf 10} and three generations of
{\bf 16} (with the exception of
the right-handed neutrinos) get their mass at $M_W$.
The Table clearly illustrates our point about the violation of the
survival principle. The last two rows contain a number of states
with masses below the corresponding scales of symmetry breaking.
The other entries of the Table are clear and conform to the
survival hypothesis. If one is weary of effective field theory
arguments, it is a straightforward exercise to obtain the masses
by direct calculation. For example the color octet states
in $(1,1,15)_A$ mix with their super heavy counterparts
in $(1,1,20)_S$; hence their see-saw-like suppressed mass.
The presence of these new states significantly alters the
unification predictions. We wish to have a rough qualitative
estimate of the new mass scales and thus it is only appropriate
to perform this at the one-loop level. It is straightforward to
write down the renormalization group equations and
solve for $M_X$, $M_C$ and $M_R$ in terms of the (unknown) unified
coupling $\alpha_U$. We find that the most interesting scenario
occurs for $M_C^2/M_X > M_R^2/M_C$ (which fixes the mass of the
octet).
We find, remarkably enough, that lowering
$M_R$ implies the lowering of $M_X$.
This is an important result which changes the nature
of the proton decay.
Clearly, the proton decay constraints set a lower limit
on $M_X$.
At the same time we make sure that the value of
$\alpha_U^{-1}$ remains perturbative.
With the scale of supersymmetry
breaking $\sim M_Z$, requiring $M_X \gtrsim 10^{15.5}$ GeV gives
\begin{equation}
M_C \gtrsim 10^{14.7}{\rm GeV} \;\; ,\quad M_R
\gtrsim 10^{13.5}{\rm GeV} \;\; ,
\end{equation}
for $\alpha_3(M_Z)=0.119$. The lower limit on $M_R$ (and $M_C$)
gets further decreased for higher $\alpha_3(M_Z)$, and for
a typical MSSM value of $\alpha_3(M_Z)=0.126$ we get
$M_C \gtrsim 10^{14.3}$GeV, $M_R \gtrsim 10^{13.0}$GeV.
Notice that the scale $M_C^2/M_X$ is actually bigger
than $M_R$. In the Table we have shown it below
only in order to separate new non-renormalizable
mass scales from the original ones.
The precise value of an intermediate scale $M_R$ should be
checked at the two-loop level including threshold effects
(however see \cite{sher89}), since it is not too far
from the unification scale. Its existence is clearly
important for neutrino masses through the see-saw,
but its impact is even more dramatic on the proton decay
predictions. If $M_R$ lies below $10^{14}$ GeV, the dominant
proton decay mode may as well be into the positron and the
neutral pion, as in the non-supersymmetric theories. This
is in sharp contradiction with the minimal supersymmetric
SU(5) theory in which this mode is highly suppressed and
dominant modes would necessarily involve kaons, not pions
(for a complete analysis and references see \cite{mura93}).
Of course, we may still have the kaon modes, however not
necessarily dominant.
Anyway, the above example illustrates nicely our point:
the combined effect of supersymmetry and internal
symmetries may lead naturally to particles with masses
much below the scale of symmetry breaking. Their impact
on unification predictions may be non-trivial and should
not be ignored. This is a dynamical question, though,
that depends on the model in question. In the next
case it will turn out, in spite of the presence of
potentially light states, that the theory chooses
the minimal SU(5) route without any intermediate scale.
\subsection{${\bf 16}$ and ${\bf \overline{16}}$ or large-scale
R-parity breaking}
Supersymmetric SO(10) models have been studied at length
with the non-renormalizable version of the see-saw mechanism\cite{sedam}.
More precisely, one chooses one (or more)
pair of ``Higgs'' in the spinorial representation ${\bf 16} $
and $\overline{{\bf 16}}$ whose VEVs induce B-L breaking and the mass
for the right-handed neutrino through the $d=4$ terms:
$m_{\nu_R} \simeq \langle {\bf 16} \overline {\bf 16}\rangle
/M_{Pl}$. This enables one to get rid of the large 126-dimensional
representations which render the high-energy behavior above
$M_X$ non asymptotically free.
The disadvantage of this program, though, is that then R-parity
is broken at a large scale $M_R$, and thus one needs additional,
ad-hoc symmetries to understand the smallness of R-parity breaking
at low energies. This appears to us more problematic than
the loss of asymptotic freedom at super high energies.
Even more important, it will turn out that, in spite of the possibly
light particles, the unification constraints push all the scales
towards $M_X$, leaving no room for intermediate scales. We still
include this example in order to show the subtlety of the issue:
it is a dynamical question whether or not one can have intermediate
mass scales.
The superpotential closely resembles the previous case's one
\begin{equation}
W=
{m_S \over 2}{\rm Tr}\, S^2+
{\lambda_S \over 3} {\rm Tr}\, S^3+
{m_A \over 2} {\rm Tr} \, A^2+
\lambda {\rm Tr}\, A^2S +
(m_\psi+\eta_A A)\psi\overline\psi\;.
\end{equation}
In complete analogy with case ({\it i}), the SO(10) symmetry forbids
self-couplings for $A$ and $\psi$. They will be generated as higher
dimensional terms, i.e. when $S$ is integrated out we get
$A^4/\langle S\rangle$, and when $A$ is integrated out one gets
$(\psi\overline\psi)^2/\langle A\rangle$. The pattern of symmetry
breaking is taken again to be (\ref{sixteen}) with $\psi$ in the
role of $\Sigma$, and the mass spectrum is obtained following a
similar reasoning. However, the LR breaking is now performed by
SU(2)$_R$ {\em doublets}, and the only state in
$\psi, \overline\psi$ not participating in the super-Higgs
mechanism is a neutral combination. The complete spectrum is now
\begin{center}
\framebox{\begin{tabular}{l|c}
\hspace{2cm}State & \hspace{1cm} Mass \\
\hline
\begin{tabular}{l}
all of $S$ in ${\bf 54}$ \\
all of $A$ in ${\bf 45}$, except $(1,1,15)_A$
\end{tabular}
& $\sim M_X$ \\
\hline
\begin{tabular}{l}
$(2,1,4)_\psi$ + $(2,1,\overline{4})_{\overline\psi}$ \\
color triplets of $(1,2,\overline{4})_\psi$ and
$(1,2,{4})_{\overline\psi}$ \\
color triplets of $(1,1,15)_A$
\end{tabular}
& $\sim M_C$ \\
\hline
\begin{tabular}{l}
one neutral and two charged combinations from \\
the color singlets of $(1,2,\overline{4})_\psi$
and $(1,2,{4})_{\overline\psi}$
\end{tabular}
& $\sim M_{R}$ \\
\hline
color octet and singlet of $(1,1,15)_A$
& $\sim Max\left[ {M_R^2\over M_C}, {M_C^2\over M_X}
\right]$ \\
\hline
\begin{tabular}{l}
neutral combination from the color singlets \\
of $(1,2,\overline{4})_\psi$ and $(1,2,{4})_{\overline\psi}$
\end{tabular}
& $\sim {M_{R}^2 \over M_X}$
\end{tabular}}
\vspace{0.5cm}
\end{center}
Only the color octet and the extra Higgs doublets could play
some role in the running. Our calculations show that the
only hope for unification is with
$M_C^2/M_X$ above $M_R$, but even in that case
the only extra
non-singlet fields below $M_R$, the heavy Higgs doublets, are
not enough to change the usual MSSM picture, and no intermediate
scales are present.
\section{Conclusions}
The minimal supersymmetric standard model, if extrapolated
to very high energies, predicts successfully the unification
of gauge couplings. This would imply a desert above $M_W$ all
the way to the unification scale at $10^{16}$ GeV, a rather
dreadful scenario. Recently, there has been great excitement
about the possibility of large compactified dimensions which
would offer plenty of new physics to fill in the desert, even
at scales as low as $1-10$ TeV \cite{extradim}.
However, it is fair to say that the unification in this
scenario \cite{ddg98} is still far from reaching the level of the
usual supersymmetric field theory \cite{mssm}, and furthermore it is
plagued by the severe proton decay problem.
Meanwhile, one is tempted to look for a possible
oasis in the desert of energies between $M_W$ and $M_X$.
We have argued here that the
supersymmetric SO(10) theory offers naturally such an
oasis. It allows for a new scale associated with
a see-saw mechanism below $10^{14}$ GeV. Admittedly,
this scale is too large to be of direct experimental
interest, but it could play an important role for
neutrino physics. Furthermore, it has a strong
effect on proton decay suggesting a possibly dominant
mode: $p \rightarrow \pi^0 e^+$. At the same time,
in this theory R-parity is exact which solves the
d=4 proton decay problem of low energy supersymmetry
and guarantees the stability of the LSP (the lightest
supersymmetric partner) \cite{amrs99}. This and other
aspects of the theory will be discussed in a
forthcoming publication.
The main reason behind the existence of a new scale below
$M_X$ is the violation of the survival principle as
we have discussed at length. In many cases supersymmetric
theories are characterized by the absence of cubic
couplings in the superpotential; this leads to an
important possibility of light states with masses of
the order $M_I^2/M$, where $M_I$
is the scale of relevant symmetry and $M$ is the cut-off
scale, as explained in the main body of this work.
Such states, which often carry exotic quantum numbers,
thus may lie at a scale much lower than $M_I$ and possibly
as low as the scales soon to be probed in new experiments.
Even if the reader is not excited by the examples we have provided,
we hope that she or he will find even more interesting
theories which will not suffer from the monotony of the
desert and which will offer new physics much closer
to the electro-weak scale.
\acknowledgments
We thank Umberto Cotti, Gia Dvali and Hossein Sarmadi
for useful discussions.
The work of B.B. is supported by the Ministry of Science and
Technology of the Republic of Slovenia; the work
of A.R. and G.S. is partially supported by EEC
under the TMR contract ERBFMRX-CT960090 and that of A.M. by
CDCHT-ULA Project No. C-898-98-05-B. A.R. is also supported by
DOE grant No. DE-FG02-97ER41036. B.B. and A.M. thank ICTP for
hospitality.
|
\section{Introduction}
The PSPC instrument onboard the ROSAT spacecraft made an 8.98 ksec
observation of the Small Magellanic Cloud on 1992 Sep 30 - Oct 2
leading to the discovery (Clark, Remillard \& Woo 1996) of RX
J0117.6-7330, a bright X-ray source within 5 arc minutes of
SMC X-1. The source was not detected in an observation of the
SMC a year earlier and was found 246 days later to be dimmer by
more than 2 orders of
magnitude. Further analysis showed that the X-ray luminosity of $2.3
\times 10^{37}$ (D/60 kpc)$^{2}$ ergs s$^{-1}$ (0.2-2.5 keV) was
derived assuming a position in the SMC (Clark, Remillard \& Woo 1997).
Spectral analysis showed the source to be relatively soft, with a
power law index of around 2.7 (although a power-law is not the best fit
model). A Fourier analysis did not reveal any significant
periodicities, with the authors lamenting an increase in spectral
noise at frequencies below 0.1 Hz.
The companion star first suggested by Clark, Remillard \& Woo (1996)
was observed by Charles, Southwell \& O'Donoghue (1996) optically
1996 January. These authors determined that the
B1-2 star of magnitude 14.2 proposed as the companion showed a strong
IR excess and Balmer lines and a reddening typical of an OB star in
the SMC, thus strengthening the association of the X-ray source with
the SMC. They also argue that the luminosity and companion type
indicates that the X-ray source is a neutron star (Coe et al. 1997).
However, Clark et al. (1997) hypothesize that the system could harbor
a black hole based on the e-folding X-ray decay time of 44 days, the
rather soft spectrum, and the lack of any neutron star rotation period
in the X-ray analysis.
We have performed a reanalysis of the
ROSAT/PSPC data and coupled it with hard X-ray observations by the
CGRO/BATSE instrument. In Section 2, we present evidence for X-ray
emission pulsed at a 22.067 second period which definitively establishes
the X-ray source as a neutron star. We show the frequency
history for RX J0117.6-7330 during this 100 day outburst which reveals
an extremely large average frequency derivative of $8.9\times10^{-11}$
Hz s$^{-1}$ corresponding to a spin-up time scale of 16 years.
The frequency derivative peaked at $1.2\times10^{-10}$ Hz
s$^{-1}$, with the pulse frequency increasing by 1.8\% during the BATSE
observations. The broad-band X-ray pulse shapes and pulsed flux are
calculated in Section 3. Section 4 summarizes our findings and
discusses RX J0117.6-7330 in the context of the Be class of HMXB's.
\section{Periodicity Search}
RX J0117.6-7330 was 5 arcminutes from the center of the PSPC
field-of-view during an 8985 second exposure taken from MJD 48895.7
- 48897.6 (MJD = JD - 2400000.5).
We determine a total source count rate of $4.43\pm0.03$
cts/sec from 0.1-2.4 keV.
Photons in a circle
of radius 1 arcminute surrounding the source position J2000 RA,Dec:
(01 17 36, -73 30 00) were extracted and barycentered using standard
FTOOLS software. A total of 33989 photons were available for timing
analysis.
These photons were collected into 5 msec bins over the full
length of the observation, a time span of 162 ksec. An FFT of the
resultant time series was then calculated, sensitive to periods in the
range from 10 msec to 81000 seconds with the power per channel
normalized to unity using the average power for all frequencies above 0.01 Hz.
No frequency derivatives were
included at this point of the analysis. A peak of power 30, normalized
as above, at a
frequency of 0.090825(2) Hz was evident in this analysis which
warranted further study despite increased noise due to the
complicated ROSAT
exposure induced window function and spacecraft wobble.
Verification of the pulsed signal comes from an archival search of
data from the BATSE instrument on the Compton Gamma-ray
Observatory. BATSE is capable of nearly continuous monitoring of hard
X-ray sources using both the earth occultation technique (Harmon et
al. 1992) and timing techniques (Bildsten et al. 1997).
A search of archival BATSE FFT results identified a
possibly related outburst some 60 days after the 1992 ROSAT
observation at a frequency of 0.0914825(2) Hz, with a frequency
derivative of $9.7(2) \times 10^{-11}$ Hz s$^{-1}$ consistent with the direction
of the SMC. With follow-up Epoch-folding based searches of the BATSE data
during the ROSAT observation, it became apparent that the originally
detected frequency was the second harmonic of the pulse frequency.
From six days of BATSE data centered on the ROSAT observation, a
barycentric pulse frequency of 0.045316682(55) Hz (MJD 48896.65) and
frequency derivative of $9.81(8)\times10^{-11}$ Hz s$^{-1}$ were determined.
Folding the ROSAT data with this frequency derivative gives a peak power
at 0.0453168(3) Hz, consistent with the BATSE pulse period.
Figure 1 shows the resultant ROSAT power spectrum calculated using the
Z$^{2}_{2}$ statistic (Buccheri et al. 1983) over a narrow frequency
range utilizing the
above-stated frequency derivative. Also plotted is the same statistic
for the BATSE data at frequencies near the ROSAT signal.
The results of searching the BATSE DISCLA channel 1 data (20-50 keV,
1.024s resolution) from 1992 August 16 (MJD 48850) to 1993 January 12
for pulsations from RX J0117-7730 are presented in Fig. 2. These
searches were performed in six day intervals, using an epoch-folding
based search (see Bildsten et al. 1997) which used only the
first and second harmonic of the pulse profile, and incorporated
a search in both pulse frequency and frequency derivative. For
intervals where pulsations were detected the pulse frequency and
frequency derivative are shown. The frequency derivative peaks at
$1.2\times 10^{-10}$ {\rm Hz~s}$^{-1}$ 25 days before the 1992 ROSAT
observations. The pulse frequency increases by 1.8\% during the
outburst. The signal is present for approximately 100 days, starting
about 34 days before the ROSAT/PSPC observation.
\section{Broad-band Pulse Profile and Flux}
Using the measured frequency and frequency derivative, we
can construct the pulse profile for RX J0117.6-7330 for both
soft and hard X-ray energies. Figure 3 shows both the ROSAT/PSPC
pulse profile for 1992 Sep 30 - Oct 2 (MJD 48895-48897)and the
CGRO/BATSE profile for MJD 48893.65 - 48899.65. Both datasets use
the pulse phase model based on the BATSE data.
The optimal ROSAT frequency and frequency derivatives are slightly
different. While the very high frequency derivative, coupled with the
long integration times (6 days for BATSE, 2 days for ROSAT) make
absolute timing comparisons slightly problematic, there is no evidence
for a loss of coherence in the BATSE folding and the pulse phases from
the two instruments should be directly comparable. Figure 3 shows the
pulse shapes for both energy ranges using an epoch at phase zero of
MJD 48896.65. An extra phase offset of 0.14 was added in order to make
the BATSE minimum correspond to phase zero.
The overall profile shapes are similar.
The peaks and minima of the lightcurves are generally in phase, but the primary
peak in the ROSAT energy range becomes the secondary peak in the BATSE
range. Similarly, primary and secondary minima are interchanged.
We have
tested our timing analysis methods using contemporaneous ROSAT/PSPC and
CGRO/BATSE observations of PSR 1509-58. In this case, we find pulse
profiles with shapes and radio-phase offsets consistent with previously
published results(Greiveldinger et al. 1995, Ulmer et al. 1993).
The overall pulse shape for RX J0117.6-7330 in both energy ranges
are similar, so it is not out of the quesion that
the shape as a whole has simply shifted.
Apparent phase shifts of simple profiles in
different energy bands have been previously observed. For example,
the 1.2-2.3 keV and 18.4-27.5 keV profiles of GS 0834-430
observed with Ginga by Akoi et al. (1992) show a complex evolution with
energy. It would be somewhat coincidental, however, for the phase shift
to be such that the minima and maxima still coincide. At this point, we
consider the peaks to be aligned, with the relative strengths to be changing.
From these pulse profiles one may determine the pulsed flux and
pulsed fraction.
Using XSPEC, we calculate a total flux of $5.1\pm0.3 \times 10^{-11}$ ergs
cm$^{-2}$ s$^{-1}$. Similar to Clark et al. 1997, we find the best fit
to the ROSAT spectrum is a combination of power-law and bremsstrahlung
or blackbody although a straight power-law fit of index $2.65\pm0.07$
is not much worse. Of the 33989 total counts extracted for the light
curve, 3829 comprise the pulsed excess giving a pulsed percentage of
$11.3\pm2.3$\%. This corresponds to a total pulsed flux in the 0.2 to
2.5 keV band of $5.6\pm1.7\times10^{-12}$ ergs cm$^{2}$ s$^{-1}$.
In the BATSE energy range, the phase-averaged pulsed fraction is more
difficult to assess. An occultation analysis of RX J0117.6-7330
detects a clear signal over the same time frame as the epoch-folding
analysis. However, source confusion could significantly contribute to
the detected flux. For a 20 day period around the time of the ROSAT
observation, the average total flux is $0.012\pm0.02$ cm$^{-2}$ s$^{-1}$.
This corresponds to an energy flux of $2.3\pm0.4 \times 10^{-10}$ ergs
cm$^{-2}$ s$^{-1}$ (assuming a power law index 3.0, 20-100 keV).
The BATSE pulsed
spectrum is best fit by a thermal bremsstrahlung model with
temperature $18 \pm 3$ keV. The integrated 20-70
keV pulsed flux is $1.8\pm0.2 \times 10^{-10}$ erg cm$^{-2}$ s$^{-1}$ ( $7.8
\times10^{37}$ (D/60 kpc)$^{2}$ ergs s$^{-1}$ ). These values provide us
with a lower limit to the pulsed fraction in the 20-70 keV range of
78\% (a lower limit since the measured occultation flux is considered
to be an upper limit to total emission due to possible source confusion
and the occultation analysis went up to 100 keV). The
directly measured flux in the 0.2-2.5 and 20-70 keV bands alone
is then at least $2.3\pm0.2 \times 10^{-10}$ erg cm$^{2}$ s$^{-1}$
($1.0\pm0.1 \times10^{38}$ (D/60 kpc)$^{2}$ ergs s$^{-1}$).
\section{Discussion}
X-ray pulsations at a period of 22.07 seconds from the bright X-ray
transient RX J0117.6-7330 have been detected in both the 0.1-2.4
and 20-70 keV energy bands. This confirms the identity of the X-ray
source as a neutron star rather than a black hole. Transient X-ray
pulsars are typically found in Be systems which is consistent with,
and supports the identification of, the proposed optical counterpart.
The CGRO/BATSE detection allows long-term monitoring of the outburst
from this source. The hard X-rays are detectable for over 100 days
starting 34 days before the ROSAT observation began. A large average
spin-up is present over the duration of the outburst resulting in a
1.8\% change in frequency. The peak frequency changes appear 10-15
days before the ROSAT observation and approximately 20 days after the
outburst began. Some of the frequency derivative may be caused
by the binary orbit. Some of the variations seen in Figure 2 near the
peak of the frequency derivative may be an orbital signature. Such
variations are weak, however, compared to the overall accretion induced
changes in frequency.
The high intrinsic spin-up rates imply that an accretion
disk is present about the neutron star, as is generally seen in the
`giant' or type I outbursts of Be/X-ray pulsars (Bildsten et al. 1997).
The measured
luminosity is at least $1.0 \times 10^{38}$ erg s$^{-1}$ during the
ROSAT observation in the combined 0.2-2.5 and 20-70 keV bands.
The peak frequency derivative was about 15\%
higher than during the ROSAT observation. However, we have no
data in the energy range from 2.5 - 20 keV. With a complicated ROSAT
X-ray spectrum and a changing pulse fraction, is difficult to
extrapolate our results to this energy range. However, it is likely
that the luminosity in this range is comparable to that measured in
the 0.2-2.5 and 20-70 keV ranges. Thus the peak luminosity, when
adjusted for higher frequency derivative and 2-20 keV emission is $\geq
1 \times 10^{38}$ erg s$^{-1}$, and was
probably higher than the conventional Eddington limit.
This is consistent with
the trend for Magellanic cloud binaries to be much brighter on average
than their galactic counterparts probably due to the absence of metals
which supply accretion inhibiting absorption
(Clark et al. 1978; van Paradijs \& McClintock 1995).
From the peak spin-up rate and standard
accretion theory (Bildsten et al. 1997 and references therein),
we can obtain a lower limit on the peak luminosity
of around $2.5\times10^{38}$ erg s$^{-1}$. This is consistent with
the mean luminosities of X-ray binaries in the SMC (van Paradijs \&
McClintock 1995). It is inconsistent with
the source being a galactic object.
The pulse profiles in the two energy bands are similar in that they both
have a double peaked structure. However, the
main and secondary peaks are interchanged.
X-ray binaries typically have pulse profiles which are often strongly
energy dependent (White, Swank \& Holt 1983). In this case, the
double peaked lightcurve in both energy bands show the same
peak-to-peak separation of 0.5.
The pulsed fraction increases from 11\% in soft X-rays to at
least 78\% in hard X-rays. If the overall morphology of the pulses is
indeed the same, with the exception of the relative strengths of the two peaks,
this may indicate a high magnetic field since at
energies above the cyclotron energy, the pulse shape is expected to
change significantly (e.g. see Sturner and Dermer 1994).
{\bf Acknowledgements }
This project made use of software and data provided by the High-Energy
Astrophysics Archival Research Center (HEASARC) located at Goddard
Space Flight Center. This work was supported at Caltech in part by NASA
NAG 5-3239.
\newpage
\section{References}
\vskip 5pt
Aoki, T., et al. 1992, PASJ, 44, 641
\vskip 5pt
Bildsten, L., et al. 1997, ApJS, 113, 367
\vskip 5pt
Buccheri, R., et al. 1983, A\&A, 128, 245
\vskip 5pt
Charles, P.A., Southwell, K.A., \& O'Donoghue, D. 1996, IAUC 6305
\vskip 5pt
Clark, G., Doxsey, R., Lie, F., Jernigan, J.G. \& van Paradijs, J. 1978, ApJ, 221, L37
\vskip 5pt
Clark, G., Remillard, R., \& Woo, J. 1996, IAUC 6282
\vskip 5pt
Clark, G., Remillard, R., \& Woo, J. 1997, ApJ, 474, L111
\vskip 5pt
Coe, M.J., Buckley, D.A.H., Charles, P.A., Southwell, K.A., \& Stevens, J. B
1998, MNRAS, 293, 43C
\vskip 5pt
Harmon, A. et al. 1993, in ``Compton Gamma-Ray Observatory'', AIP
Conf. Proceedings 280, (AIP: New York), 314
\vskip 5pt
Greiveldinger, C., Caucino, S., Massaglia, S., Ogelman, H. \& Trussoni,
E. 1995, ApJ, 454, 855
\vskip 5pt
Sturner, S.J. \& Dermer, C.D. 1994, A\&A, 284, 161
\vskip 5pt
Ulmer, M.P., et al. 1993, ApJ, 417, 738
\vskip 5pt
van Paradijs, J. \& McClintock, J.E. 1995, in "X-ray binaries", ed.
Lewin, W.H.G., van Paradijs, \& van den Heuvel, E.P.J., Cambridge
University Press, p. 113
\vskip 5pt
White, N.E., Swank, J.H., \& Holt, S.S. 1983, ApJ, 270, 711
\vskip 5pt
White, N.E., Giommi, P. \& Angelini, L. 1995, BAAS, 185
\newpage
\centerline{\psfig{figure=rxj0117_figure1.ps}}
{\bf Figure 1}: The ROSAT power distribution encompassing the first
harmonic of the pulsed frequency. The BATSE distribution is for a narrow
range around the ROSAT detection frequency.
\vskip 12pt
\newpage
\centerline{\psfig{figure=rxj0117_figure2.ps}}
{\bf Figure 2}: BATSE pulse timing analysis of RX J0117.6-7330. The
top panel is the frequency history and the bottom panel the frequency
rate history. Both plots use the best fit values for 6 day intervals
with the ROSAT observation date shown by the dashed line.
\vskip 12pt
\newpage
centerline{\psfig{figure=rxj0117_figure3.ps}}
{\bf Figure 3}: The folded light curves for ROSAT 0.1-2.4 keV (top)
and BATSE 20 - 70 keV (bottom) data. The BATSE profile is limited to six
Fourier harmonics, resulting in the smooth shape. Flux errors
are given for a set of approximately independent phases.
\end{document}
|
\subsection*{Acknowledgments}
The author would like to thank M.\ Bando and T.\ Noguchi for useful
discussions and careful reading of the manuscript. This work was
supported in part by the Grant-in-Aid for JSPS Research Fellowships.
|
\section{Bethe ansatz for the Hubbard model}
\subsection{Eigenfunctions and eigenvalues}
The Hamiltonian of the one-dimensional Hubbard model on a periodic
$L$-site chain may be written as
\beq \label{ham}
H = - \sum_{j=1}^L \sum_{\s = \auf, \ab}
(c_{j, \s}^+ c_{j+1, \s} + c_{j+1, \s}^+ c_{j, \s})
+ U \sum_{j=1}^L
(n_{j \auf} - \tst{\2})(n_{j \ab} - \tst{\2}) \qd.
\eeq
$c_{j, \s}^+$ and $c_{j, \s}$ are creation and annihilation operators
of electrons in Wannier states, and periodicity is guaranteed by
setting $c_{L+1, \s} = c_{1, \s}$. $n_{j,\s} = c_{j, \s}^+ c_{j, \s}$
is the particle number operator for electrons of spin $\sigma$ at site
$j$, $U$ is the coupling constant. The eigenvalue problem for the
Hubbard Hamiltonian (\ref{ham}) was solved by Lieb and Wu \cite{LiWu68}
using the nested Bethe ansatz \cite{Yang67}. The Hubbard Hamiltonian
conserves the number of electrons $N$ and the number of down spins $M$.
The corresponding Schr\"odinger equation can therefore be solved for
fixed $N$ and $M$. Since the Hamiltonian is invariant under
particle-hole transformations and under reversal of spins
\cite{LiWu68}, we may set $2M \le N \le L$. We shall denote the
positions and spins of the electrons by $x_j$ and $\s_j$, respectively.
The Bethe ansatz eigenfunctions of the Hubbard Hamiltonian (\ref{ham})
depend on the relative ordering of the $x_j$. There are $N!$ possible
orderings of the coordinates of $N$ electrons. Any ordering may be
related to a permutation $Q$ of the numbers $1, \dots, N$ through the
inequality
\beq \label{sectorq}
1 \le x_{Q1} \le x_{Q2} \le \dots \le x_{QN} \le L \qd.
\eeq
This inequality divides the configuration space of $N$ electrons into
$N!$ sectors, which can be labeled by the permutations~$Q$. The Bethe
ansatz eigenfunctions of the Hubbard Hamiltonian (\ref{ham}) in the
sector $Q$ are given as
\beq \label{wwf}
\ps (x_1, \dots, x_N; \s_1, \dots, \s_N) =
\sum_{P \in S_N} \sign(PQ) \, \ph_P (\s_{Q1}, \dots, \s_{QN})
\exp \left( \i \sum_{j=1}^N k_{Pj} x_{Qj} \right) \qd.
\eeq
Here the $P$-summation extends over all permutations of the numbers
$1, \dots, N$. These permutations form the symmetric group $S_N$.
The function $\sign(Q)$ is the sign function on the symmetric group,
which is $- 1$ for odd permutations and $+ 1$ for even permutations.
The spin dependent amplitudes $\ph_P (\s_{Q1}, \dots, \s_{QN})$ can be
found in Woynarovich's paper \cite{Woynarovich82a}. They are of the
form of the Bethe ansatz wave functions of an inhomogeneous XXX spin
chain,
\beq \label{wswf}
\ph_P (\s_{Q1}, \dots, \s_{QN}) = \sum_{\p \in S_M}
A(\la_{\p 1}, \dots, \la_{\p M})
\prod_{l=1}^M F_P (\la_{\p l}; y_l) \qd.
\eeq
Here $F_P (\la; y)$ is defined as
\beq
F_P (\la; y) = \frac{1}{\la - \sin k_{Py} + \i U/4}
\prod_{j=1}^{y-1} \frac{\la - \sin k_{Pj} - \i U/4}
{\la - \sin k_{Pj} + \i U/4} \qd,
\eeq
and the amplitudes $A(\la_1, \dots, \la_M)$ are given by
\beq \label{wwfsa}
A(\la_1, \dots, \la_M) = \prod_{1 \le m < n \le M}
\frac{\la_m - \la_n - \i U/2}{\la_m - \la_n} \qd.
\eeq
$y_j$ in the above equations denotes the position of the $j$th down
spin in the sequence $\s_{Q1}, \dots, \s_{QN}$. The $y$'s are thus
`coordinates of down spins on electrons'. Below we shall illustrate
the notation through an explicit example.
The wave functions (\ref{wwf}) are characterized by two sets of
quantum numbers $\{k_j\}$ and $\{\la_l\}$. These quantum numbers
may be generally complex. The $k_j$ and $\la_l$ are called charge
momenta and spin rapidities, respectively. The charge momenta and spin
rapidities satisfy the Lieb-Wu equations
\bea \label{bak}
e^{\i k_j L} & = & \prod_{l=1}^M \frac{\la_l - \sin k_j - \i U/4}
{\la_l - \sin k_j + \i U/4} \qd,
\qd j = 1, \dots, N \qd, \\
\label{bas}
\prod_{j=1}^N \frac{\la_l - \sin k_j - \i U/4}
{\la_l - \sin k_j + \i U/4} & = &
\prod_{m=1 \atop m \ne l}^M \frac{\la_l - \la_m - \i U/2}
{\la_l - \la_m + \i U/2} \qd,
\qd l = 1, \dots, M \qd.
\eea
A derivation of the wave function (\ref{wwf}) and the Lieb-Wu equations
(\ref{bak}), (\ref{bas}) is presented in appendices A and B.
The wave functions (\ref{wwf}) are joint eigenfunctions of the
Hubbard Hamiltonian (\ref{ham}) and the momentum operator%
\footnote{For a proper definition of the momentum operator see
appendix B of \cite{GoMu97b}} with eigenvalues
\beq \label{enmom}
E = - 2 \sum_{j=1}^N \cos k_j + \frac{U}{4}(L - 2N) \qd, \qd
P = \left( \sum_{j=1}^N k_j \right) \mod \, 2\p \qd.
\eeq
The `coordinates of down spins' $y_j$ which enter (\ref{wswf}) depend
on $(\s_1, \dots, \s_N)$ {\it and} on $(x_1, \dots, x_N)$. The
following example should help to understand the notation. Let $L = 12$,
$N = 5$, $M = 2$, and let, for example, $(x_1, \dots, x_5) =
(7,3,5,1,8)$, $(\s_1, \dots, \s_5) = (\auf \auf \auf \ab \ab)$. Then
$x_4 \le x_2 \le x_3 \le x_1 \le x_5$, i.e.\ $Q = (4,2,3,1,5)$. It
follows that $(x_{Q1}, \dots, x_{Q5}) = (1,3,5,7,8)$ and $(\s_{Q1},
\dots, \s_{Q5}) = (\ab \auf \auf \auf \ab)$. Thus $y_1 = 1$, $y_2 = 5$.
Whenever it will be necessary, we shall indicate the dependence of the
wave functions (\ref{wwf}) on the charge momenta and spin rapidities
by subscripts, $\ps = \ps_{k_1, \dots, k_N; \la_1, \dots, \la_M}$.
Let us consider the symmetries of the eigenfunctions under permutations,
\bea \label{perm1}
\ps (x_{P1}, \dots, x_{PN}; \s_{P1}, \dots, \s_{PN}) & = &
\sign(P) \ps (x_1, \dots, x_N; \s_1, \dots, \s_N) \qd, \qd
P \in S_N \qd, \\ \label{perm2}
\ps_{k_{P1}, \dots, k_{PN}; \la_1, \dots, \la_M} & = &
\sign(P) \ps_{k_1, \dots, k_N; \la_1, \dots, \la_M} \qd, \qd
P \in S_N \qd, \\ \label{perm3}
\ps_{k_1, \dots, k_N; \la_{P1}, \dots, \la_{PM}} & = &
\ps_{k_1, \dots, k_N; \la_1, \dots, \la_M} \qd, \qd
P \in S_M \qd.
\eea
Equation (\ref{perm1}) means that the eigenfunctions respect the Pauli
principle. (\ref{perm2}) and (\ref{perm3}) describe their properties
with respect to permutations of the quantum numbers. They are totally
antisymmetric with respect to interchange of the charge momenta $k_j$,
and they are totally symmetric with respect to interchange of the spin
rapidities $\la_l$. Hence, in order to find all Bethe ansatz wave
functions we have to solve the Lieb-Wu equations (\ref{bak}),
(\ref{bas}) modulo permutations of the sets $\{k_j\}$ and $\{\la_l\}$.
The $k_j$'s have to be mutually distinct, since otherwise the wave
function vanishes due to (\ref{perm2}). In fact, the $\la_l$'s have to
be mutually distinct, too. This is called the `Pauli principle for
interacting Bosons' (see \cite{KBIBo}). We would like to emphasize that
there are no further restrictions on the solutions of (\ref{bak}),
(\ref{bas}). In particular, the spin and charge rapidities do {\it not}
have to be real.
Bethe ansatz states on a finite lattice of length $L$ that have finite
momenta $k_j$ and rapidities $\la_l$, a non-negative value of the
total spin ($N - 2M > 0$), and a total number of electrons not larger
than the length of the lattice ($N \le L$) are called {\it regular}
(cf.\ \cite{EKS92a}, page 562).
There exist two discrete symmetries of the model which can be used to
obtain additional eigenstates from the regular ones \cite{LiWu68}.
The Hamiltonian is invariant under exchange of up and down spins.
This symmetry allows for obtaining eigenstates with negative value
$N- 2M$ of the total spin from eigenstates with positive value of the
total spin. This symmetry does not affect the number of electrons.
Thus, its action on regular states does not lead above half filling.
States above half filling ($N > L$) can be obtained by employing the
transformation $c_{j \s} \rightarrow (- 1)^j c_{j \s}^+$, $c_{j \s}^+
\rightarrow (- 1)^j c_{j \s}$, $\s = \auf, \ab$, which leaves the
Hamiltonian (\ref{ham}) invariant, but maps the empty Fock state
$|0\>$ to the completely filled Fock state $|\auf \ab\>$.
\subsection{SO(4) symmetry}
The Hubbard Hamiltonian (\ref{ham}) is invariant under rotations in
spin space. The corresponding su(2) Lie algebra is generated by the
operators
\beq \label{rot}
\begin{array}{r@{\qd, \qd}c@{\qd, \qd}l}
\dst{\z = \sum_{j=1}^L c_{j \auf}^+ c_{j \ab}} &
\dst{\z^\dagger = \sum_{j=1}^L c_{j \ab}^+ c_{j \auf}} &
\dst{\z^z = \tst{\2} \sum_{j=1}^L (n_{j \ab} - n_{j \auf}) \qd.}
\\[2ex]
[\z,\z^\dagger] = - 2 \z^z &
[\z,\z^z] = \z &
[\z^\dagger,\z^z] = - \z^\dagger \qd.
\end{array}
\eeq
For lattices of even length $L$ there is another representation of
su(2), which commutes with the Hubbard Hamiltonian \cite{HeLi71,%
Yang89,Pernici90}.
This representation generates the so-called $\h$-pairing symmetry,
\beq \label{eta}
\begin{array}{r@{\qd, \qd}c@{\qd, \qd}l}
\dst{\h = \sum_{j=1}^L (- 1)^j c_{j \auf} c_{j \ab}} &
\dst{\h^\dagger = \sum_{j=1}^L (- 1)^j
c_{j \ab}^+ c_{j \auf}^+} &
\dst{\h^z = \tst{\2} \sum_{j=1}^L (n_{j \ab} + n_{j \auf})
- \tst{\2 L} \qd.}
\\[2ex]
[\h,\h^\dagger] = - 2 \h^z &
[\h,\h^z] = \h &
[\h^\dagger,\h^z] = - \h^\dagger \qd.
\end{array}
\eeq
The generators of both algebras commute with one-another. They combine
into a representation of su(2)$\oplus$su(2).
The $\h$-pairing symmetry connects sectors of the Hilbert space with
different numbers of electrons. The operator $\h^\dagger$, for instance,
creates a local pair of electrons of opposite spin and momentum $\p$.
Hence, in order to consider the action of the $\h$-symmetry on
eigenstates we write them in second quantized form.
\beq \label{sq}
|k_1, \dots, k_N; \la_1, \dots, \la_M \> =
\sum_{x_1, \dots, x_N = 1}^L
\ps_{k_1, \dots, k_N; \la_1, \dots, \la_M}
(x_1, \dots, x_N; \s_1, \dots, \s_N)
c_{x_1, \s_1}^+ \dots c_{x_N, \s_N}^+ |0\> \qd,
\eeq
where $\s_1 = \dots = \s_M = \ab$ and $\s_{M + 1} = \dots = \s_N =
\auf$. It is easily seen that
\beq
(\z^z + \h^z) |k_1, \dots, k_N; \la_1, \dots, \la_M \> =
(M - \tst{\frac{L}{2}})
|k_1, \dots, k_N; \la_1, \dots, \la_M \> \qd.
\eeq
Here $M - \frac{L}{2}$ is integer, since $L$ is even. Therefore the
symmetry group generated by the representations (\ref{rot}), (\ref{eta})
is SO(4) rather than SU(2)$\times$SU(2) \cite{YaZh90}.
It was shown in \cite{EKS92a} that the {\it regular} Bethe ansatz
states are lowest weight vectors of both su(2) symmetries (\ref{rot})
and
(\ref{eta}),
\beq \label{lw}
\z |k_1, \dots, k_N; \la_1, \dots, \la_M \> = 0 \qd, \qd
\h |k_1, \dots, k_N; \la_1, \dots, \la_M \> = 0 \qd.
\eeq
This is an important theorem. It was the prerequsite for the proof
of completeness (see section D) of the Bethe ansatz for the Hubbard
model in \cite{EKS92b}. The proof of (\ref{lw}) is direct but lengthy
\cite{EKS92a}. $\z$ and $\h$ are applied to the states (\ref{sq}), and
the Lieb-Wu equations (\ref{bak}), (\ref{bas}) are used to reduce the
resulting expressions to zero. We would like to emphasize that the
proof of (\ref{lw}) is {\it not} restricted to real solutions of the
Lieb-Wu equations. It goes through for all solutions corresponding to
regular Bethe ansatz states including the strings.
Since the two su(2) symmetries (\ref{rot}), (\ref{eta}) leave the
Hubbard Hamiltonian (\ref{ham}) invariant, additional eigenstates which
do not belong to the regular Bethe ansatz can be obtained by applying
$\z^\dagger$ and $\h^\dagger$ to regular Bethe ansatz eigenstates. Since
\bea
\z^z |k_1, \dots, k_N; \la_1, \dots, \la_M \> & = &
(M - \tst{\frac{N}{2}}) |k_1, \dots, k_N; \la_1, \dots, \la_M \>
\qd, \\
\h^z |k_1, \dots, k_N; \la_1, \dots, \la_M \> & = &
\tst{\2}(N - L) |k_1, \dots, k_N; \la_1, \dots, \la_M \> \qd,
\eea
a state $|k_1, \dots, k_N; \la_1, \dots, \la_M \>$ has spin
$\2 (N - 2M)$ and $\h$-spin $\2 (L - N)$. The dimension of the
corresponding multiplet is thus given by
\beq \label{so4mult}
\dim_{M, N} = (N - 2M + 1)(L - N + 1) \qd.
\eeq
The states in this multiplet are of the form
\beq \label{so4ext}
|k_1, \dots, k_N; \la_1, \dots, \la_M; \a; \begin{equation} \> =
(\z^\dagger)^\a (\h^\dagger)^\begin{equation}
|k_1, \dots, k_N; \la_1, \dots, \la_M \> \qd,
\eeq
where $\a = 0, \dots, N - 2M$ and $\begin{equation} = 0, \dots, L - N$. Note that
states of the form (\ref{so4ext}) can be obtained from regular Bethe
ansatz states with $\tilde N \ge N$, $\tilde M \ge M$ by formally
setting some of the charge momenta and spin rapidities equal to
infinity \cite{Woynarovich82a,Gaudin83,FaTa84}.
\subsection{Discrete Takahashi equations}
Let us now formulate Takahashi's string hypothesis \cite{Takahashi72}
more precisely: All regular solutions $\{k_j\}$, $\{\la_l\}$ of the
Lieb-Wu equations (\ref{bak}), (\ref{bas}) consist of three different
kinds of configurations.
\begin{enumerate}
\item
A single real momentum $k_j$.
\item
$m$ $\la$'s combining into a $\La$ string. This includes the case
$m = 1$, which is just a single $\La_\a$.
\item
$2m$ $k$'s and $m$ $\la$'s combining into a $k$-$\La$ string.
\end{enumerate}
For large lattices ($L \gg 1$) and a large number of electrons
($N \gg 1$), almost all strings are close to ideal, i.e.\ the imaginary
parts of the $k$'s and $\la$'s are almost equally spaced.
For ideal $\La$ strings of length $m$ the rapidities involved are
\beq \label{ideal1}
\La_\a^{m, j} = \La_\a^m + (m - 2j +1) \tst{\frac{\i U}{4}} \qd.
\eeq
Here $\a$ enumerates the strings of the same length $m$, and $j = 1,
\dots, m$ counts the $\la$'s involved in the $\a$th $\La$ string of
length $m$. $\La_\a^m$ is the real center of the string.
The $k$'s and the $\la$'s involved in an ideal $k$-$\La$ string are
(for $U > 0$)
\bea
k_\a^1 & = & \p - \arcsin({\La'}_\a^m + m \tst{\frac{\i U}{4}})
\qd, \nonumber\\ \\
k_\a^2 & = & \arcsin({\La'}_\a^m + (m - 2) \tst{\frac{\i U}{4}})
\qd, \nonumber\\ \\
k_\a^3 & = & \p - k_\a^2 \qd, \nonumber\\ \\
& \vdots & \label{idealk} \\
k_\a^{2m - 2} & = & \arcsin({\La'}_\a^m - (m - 2)
\tst{\frac{\i U}{4}}) \qd, \nonumber\\ \\
k_\a^{2m - 1} & = & \p - k_\a^{2m - 2} \qd, \nonumber\\ \\
k_\a^{2m} & = & \p - \arcsin({\La'}_\a^m - m \tst{\frac{\i U}{4}})
\qd, \nonumber\\
\eea
and
\beq \label{ideal3}
{\La'}_\a^{m, j} = {\La'}_\a^m + (m - 2j + 1) \tst{\frac{\i U}{4}}
\qd.
\eeq
Again $m$ denotes the `length' of the string, $\a$ enumerates strings
of length $m$, and $j$ counts the $\la$'s involved in a given string.
${\La'}_\a^m$ is the real center of the $k$-$\La$ string. The branch
of $\arcsin(x)$ in (\ref{idealk}) is fixed as $- \pi/2 \le \text{Re}
(\arcsin (x)) \le \pi/2$.
The string hypothesis assumes that almost all solutions of the Lieb-Wu
equations (\ref{bak}), (\ref{bas}) are approximately given by
(\ref{ideal1})-(\ref{ideal3}) with exponentially small corrections of
order $\CO (\exp( - \de L))$, where $\de$ is real and positive and
depends on the specific string under consideration.
Using the string hypothesis inside the Lieb-Wu equations (\ref{bak}),
(\ref{bas}) and taking logarithms afterwards, we arrive at the
following form of Bethe ansatz equations for strings, which we call
discrete Takahashi equations
\begin{eqnarray} \label{t1}
k_j L & = & 2 \pi I_j - \sum_{n=1}^\infty \sum_{\alpha = 1}^{M_n}
\theta \left(
\frac{\sin k_j - \Lambda_\alpha^n}{nU/4} \right)
- \sum_{n=1}^\infty \sum_{\alpha = 1}^{M_n'}
\theta \left(
\frac{\sin k_j - {\Lambda'}_\alpha^n}{nU/4} \right),
\\ \label{t2}
\sum_{j=1}^{N - 2M'} \theta \left(
\frac{\Lambda_\alpha^n - \sin k_j}{nU/4} \right) & = &
2 \pi J_\alpha^n +
\sum_{m=1}^\infty \sum_{\beta = 1}^{M_m}
\Theta_{nm} \left(
\frac{\Lambda_\alpha^n - \Lambda_\beta^m}{U/4} \right),
\\ \label{t3}
L [\arcsin({\Lambda'}_\alpha^n + n \tst{\frac{\i U}{4}})
+ \arcsin({\Lambda'}_\alpha^n - n \tst{\frac{\i U}{4}})] & = &
2 \pi {J'}_\alpha^n +
\sum_{j=1}^{N - 2M'} \theta \left(
\frac{{\Lambda'}_\alpha^n - \sin k_j}{nU/4} \right) +
\sum_{m=1}^\infty \sum_{\beta = 1}^{M_m'}
\Theta_{nm} \left(
\frac{{\Lambda'}_\alpha^n - {\Lambda'}_\beta^m}{U/4}
\right).
\end{eqnarray}
Here we assumed $L$ to be even. $I_j$, $J_\alpha^n$, and
${J'}_\alpha^n$ are integer or half-odd integer numbers, according to
the following prescriptions: $I_j$ is integer (half odd integer), if
$\sum_m (M_m + M_m')$ is even (odd); the $J_\a^n$ are integer (half odd
integer), if $N - M_n$ is odd (even); the ${J'}_\a^n$ are integer (half
odd integer), if $L - (N - M_n')$ is odd (even). $M_n$ and $M_m'$ are
the numbers of $\Lambda$ strings of length $n$, and $k$-$\Lambda$
strings of length $m$ in a specific solution of the system (\ref{t1})-%
(\ref{t3}). $M' = \sum_{n=1}^\infty n M_n'$, is the total number of
$\la$'s involved in $k$-$\La$ strings. The integer (half-odd integer)
numbers in (\ref{t1})-(\ref{t3}) have ranges
\begin{eqnarray} \label{r1}
&& - \frac{L}{2} < I_j \le \frac{L}{2}, \\ \label{r2}
&& |J_\alpha^n| \le \frac{1}{2}
\left(N - 2M' - \sum_{m=1}^\infty t_{nm} M_m - 1 \right), \\
\label{r3}
&& |{J'}_\alpha^n| \le \frac{1}{2}
\left(L - N + 2M' - \sum_{m=1}^\infty t_{nm} M_m' - 1 \right),
\end{eqnarray}
where $t_{mn} = 2 \min (m,n) - \delta_{mn}$. The functions $\th$ and
$\Theta_{nm}$ in (\ref{t1})-(\ref{t3}) are defined as $\theta(x) =
2 \arctan(x)$, and
\begin{equation} \label{defthetas}
\Theta_{nm} (x) = \left\{ \begin{array}{l}
{\displaystyle
\theta \left( \frac{x}{|n - m|} \right) +
2 \theta \left( \frac{x}{|n - m| + 2} \right) + \cdots +
2 \theta \left( \frac{x}{n + m - 2} \right) +
\theta \left( \frac{x}{n + m} \right), \: \text{if} \quad
n \ne m,} \\[3ex]
{\displaystyle
2 \theta \left( \frac{x}{2} \right) +
2 \theta \left( \frac{x}{4} \right) + \cdots +
2 \theta \left( \frac{x}{2n - 2} \right) +
\theta \left( \frac{x}{2n} \right), \: \text{if} \quad n = m.}
\end{array} \right.
\end{equation}
In terms of the parameters of the ideal strings total energy and
momentum (\ref{enmom}) are expressed as
\bea \label{mom}
P & = & \left[ \sum_{j=1}^{N - 2M'} k_j -
\sum_{n=1}^\infty \sum_{\a = 1}^{M_n'}
\left(2 \, \Re \arcsin \left( {\La'}_\a^n +
n \tst{\frac{\i U}{4}} \right) -
(n + 1) \p \right) \right] \mod 2 \p \qd, \\
\label{en}
E & = & - 2 \sum_{j=1}^{N - 2M'} \cos(k_j) +
4 \sum_{n=1}^\infty \sum_{\a = 1}^{M_n'}
\Re \sqrt{1 - \left( {\La'}_\a^n +
n \tst{\frac{\i U}{4}} \right)^2}
+ \frac{U}{4} (L - 2N) \qd.
\eea
Equations (\ref{t1})-(\ref{r3}) can be used to study all excitations
of the Hubbard model in the thermodynamic limit. They are the basis
for the derivation of Takahashi's integral equations \cite{Takahashi72},
which determine the thermodynamics of the Hubbard model (see sections
VI and VII). Applications of (\ref{t1})-(\ref{r3}) are usually based
on the following assumptions.
\begin{enumerate}
\item
Any set of non-repeating (half odd) integers $I_j$, $J_\a^n$,
${J'}_\a^n$ subject to the constraints (\ref{r1})-(\ref{r3}) specifies
one and only one solution $\{k_j\}$, $\{\La_\a^n\}$, $\{{\La'}_\a^n\}$
of equations (\ref{t1})-(\ref{t3}).
\item
The solutions $\{k_j\}$, $\{\La_\a^n\}$, $\{{\La'}_\a^n\}$ of
(\ref{t1})-(\ref{t3}) specified by a set of non-repeating (half odd)
integers $I_j$, $J_\a^n$, ${J'}_\a^n$ subject to (\ref{r1})-(\ref{r3})
are in one-to-one correspondence to solutions of the Lieb-Wu equations
(\ref{bak}), (\ref{bas}).
\item
For large $L$ and $N$ almost every solution $\{k_j\}$, $\{\la_l\}$ of
the Lieb-Wu equations (\ref{bak}), (\ref{bas}) is exponentially close to
the corresponding solution $\{k_j\}$, $\{\La_\a^n\}$, $\{{\La'}_\a^n\}$
of the discrete Takahashi equations, which means that the strings
contained in $\{k_j\}$, $\{\la_l\}$ are well approximated by the ideal
strings determined by $\{k_j\}$, $\{\La_\a^n\}$, $\{{\La'}_\a^n\}$.
\end{enumerate}
\subsection{Completeness of the Bethe ansatz}
The proof of completeness of the Bethe ansatz given in \cite{EKS92b}
is based on assumptions (i) and (ii) above. Similar assumptions were
proved for other Bethe ansatz solvable models \cite{KBIBo}. Note that
assumption (ii) does {\it not} mean that the classification of the
solutions of the Lieb-Wu equations (\ref{bak}), (\ref{bas}) into strings
is actually given by (\ref{t1})-(\ref{r3}). There may be a
redistribution between different kinds of strings. This phenomenon was
observed in a number of Bethe ansatz solvable models and was carefully
studied by examples \cite{Bethe31,EKS92c,EKS92b} (see also section
V.B). It turned out that the redistribution did in no case affect the
total number of solutions of the Bethe ansatz equations.
Using (i) and (ii) above, the proof of completeness reduces to a
combinatorial problem based on (\ref{r1})-(\ref{r3}) \cite{EKS92b}.
From (\ref{r1})-(\ref{r3}) we read off the numbers of allowed values
of the (half odd) integers $I_j$, $J_\a^n$, ${J'}_\a^n$ in a given
configuration $\{M_n\}$, $\{M_n'\}$ of strings. These numbers are
\begin{enumerate}
\item
$L$ for a free $k_j$ (not involved in a $k$-$\La$ string),
\item
$N - 2M' - \sum_{m=1}^\infty t_{nm} M_m$ for a $\La$ string of length
$n$,
\item
$L - N + 2M' - \sum_{m=1}^\infty t_{nm} M_m'$ for a $k$-$\La$ string
of length $n$.
\end{enumerate}
The total number of ways to select the $I_j$, $J_\a^n$, ${J'}_\a^n$
(recall that they are assumed to be non-repeating) for a given
configuration $\{M_n\}$, $\{M_n'\}$ is thus
\beq
n(\{M_n\},\{{M'}_n\}) = {L \choose N - 2M'}
\prod_{n=1}^\infty
{N - 2M' - \sum_{m=1}^\infty t_{nm} M_m \choose M_n}
\prod_{n=1}^\infty
{L - N + 2M' - \sum_{m=1}^\infty t_{nm} M_m' \choose M_n'} \qd.
\eeq
Hence, the number of regular Bethe ansatz states for given numbers $N$
of electrons and $M$ of down spins is
\beq \label{nreg}
n_{\rm reg} (M,N) = \sum_{\{M_n\}, \{M_n'\}} n(\{M_n\},\{M_n'\})
\qd,
\eeq
where the summation is over all configurations of strings which satisfy
the constraints $N - 2M' \ge 0$ and $M = \sum_{m=1}^\infty m(M_m +
M_m')$. Finally, the total number of states (\ref{so4ext}) in the SO(4)
extended Bethe ansatz is
\beq \label{ntot}
n_{\rm tot} (L) = \sum_{M, N} n_{\rm reg} (M, N) \dim_{M, N}
= \sum_{M, N} n_{\rm reg} (M, N)
(N - 2M + 1)(L - N +1) \qd,
\eeq
where the sum is over all $M$, $N$ with $0 \le 2M \le N \le L$. The
sums (\ref{nreg}) and (\ref{ntot}) were calculated in \cite{EKS92b}.
It turns out that
\beq
n_{\rm tot} (L) = 4^L \qd,
\eeq
which is the dimension of the Hilbert space of the Hubbard model on an
$L$-site chain.
Let us list again the essential steps that led to the above proof of
completeness:
\begin{enumerate}
\item
Impose periodic boundary conditions.
\item
Take Woynarovich's wave function (\ref{wwf})-(\ref{wwfsa}).
\item
Define the Bethe ansatz in the narrow sense of regular Bethe ansatz
(see below (\ref{perm3})). This eliminates infinite $k$'s and $\la$'s
whose multiplicities are not under control.
\item
Prove the lowest weight theorem (\ref{lw}). Then gluing back solutions
with infinite $k$'s and $\la$'s is equivalent to considering the
multiplets (\ref{so4ext}).
\item
The multiplicities of occupation of infinite $k$'s and $\la$'s are given
by the dimensions (\ref{so4mult}) of the multiplets (\ref{so4ext}).
\item
Use Takahashi's integers (\ref{r1})-(\ref{r3}) for counting of the
regular Bethe ansatz states.
\end{enumerate}
\section{Lieb-Wu equations for a single down spin (I) -- Graphical
solution}
In this section we study the Lieb-Wu equations (\ref{bak}), (\ref{bas})
in the most simple non-trivial case, when there is only one down spin,
$M = 1$. For pedagogical clarity some emphasis will be on the most
instructive cases $N = 2$ and $N = 3$. These are the cases which we
also studied numerically (cf.\ section V). Some of the analytical
calculations, however, are presented for general $N$, simply because
the general arguments are simple enough and enable treating the
cases $N = 2$ and $N = 3$ to some extent simultaneously.
The Lieb-Wu equations for $N = 2$ and $M = 1$ were studied before in
appendix B of \cite{EKS92b}. There the emphasis was on the
redistribution phenomenon mentioned in section II.D. For $U = 0$ the
Hubbard Hamiltonian turns into a free tight binding Hamiltonian,
and there is no bound state of electrons ($k$-$\La$ string) left.
It is therefore clear that bound states decay as the coupling becomes
weaker. In appendix B of \cite{EKS92b} it was shown that each time
a $k$-$\La$ string disappears from the spectrum at a certain critical
value of the coupling $U > 0$, a new real solution emerges. Here we take
a slightly different point of view. We fix $U$ and study the solutions
for large finite $L$. It turns out that there is no redistribution
phenomenon for the most simple $k$-$\La$ strings consisting of two
complex conjugated $k$'s and one real $\la$ as $L \rightarrow \infty$.
These strings always exist for large enough finite $L$, and their
number is in accordance with the counting implied by Takahashi's
discrete equations (\ref{t1})-(\ref{r3}). In this respect the $k$-$\La$
strings of the Hubbard model are different from the $\La$ strings in
the XXX spin chain \cite{EKS92c}.
For $M = 1$ the Lieb-Wu equations (\ref{bak}), (\ref{bas}) read
\bea \label{bak1}
&& e^{\i k_j L} = \frac{\La - \sin k_j - \i U/4}
{\La - \sin k_j + \i U/4} \qd, \qd
j = 1, \dots, N \qd,\\ \label{bas1}
&& \prod_{j = 1}^N \frac{\La - \sin k_j - \i U/4}
{\La - \sin k_j + \i U/4} = 1 \qd.
\eea
Equation (\ref{bas1}) can be replaced by the equation for the
conservation of momentum,
\beq \label{emom}
e^{\i (k_1 + \dots + k_N)L} = 1 \qd.
\eeq
(\ref{bak1}) and (\ref{emom}) follow from (\ref{bak1}) and (\ref{bas1})
and vice versa. Let us take the logarithm of (\ref{emom}) and solve
(\ref{bak1}) for $\sin k_j - \La$. We obtain the equations
\bea \label{momlog}
(k_1 + \dots + k_N) \mod \, 2 \p & = & \frac{m2\p}{L} \qd, \qd
m = 0, \dots, L - 1 \qd, \\ \label{laofk}
\sin k_j - \La & = & \frac{U}{4} \ctg \left( \frac{k_j L}{2}
\right)
\qd, \qd j = 1, \dots, N \qd,
\eea
which are equivalent to (\ref{bak1}) and (\ref{bas1}) but more
convenient for the further discussion.
\subsection{\boldmath All charge momenta $k_j$ real}
The equation
\beq \label{laofq}
\sin q - \La = \frac{U}{4} \ctg \left( \frac{q L}{2}
\right)
\eeq
is easily solved graphically for $q$ as a function of $\La$ (see
figure 1). It has at least $L$ branches of solutions $q_\ell (\La)$
belonging to the interval $0 \le q < 2 \p$. There is at least one
branch with $(\ell - 1) \frac{2\p}{L} < q < \ell \frac{2\p}{L}$. Yet,
for $\frac{\p}{2} \le q \le \frac{3 \p}{2}$ there may be more than
one solution in the interval $[(\ell - 1) \frac{2\p}{L}, \ell
\frac{2\p}{L}]$, if $U$ or $L$ is too small. We have the following
uniqueness condition,
\beq
- 1 = \min_{0 \le q < 2 \p} \6_q \sin q >
\max_{0 \le q < 2 \p} \6_q \frac{U}{4}
\ctg \left( \frac{qL}{2} \right) =
\max_{0 \le q < 2 \p} - \frac{UL}{8}
\sin^{-2} \left( \frac{qL}{2} \right) = - \frac{UL}{8} \qd,
\eeq
which is equivalent to
\beq \label{tcd}
L > \frac{8}{U} \qd.
\eeq
We call this condition Takahashi condition. In the following we will
assume the Takahashi condition to hold.
Equation (\ref{laofq}) defines $\La$ as a function of $q$. We have
\beq
\frac{d \La}{dq} = \cos q + \frac{UL}{8} \sin^{-2} \left(
\frac{qL}{2} \right) \qd.
\eeq
Then, using (\ref{tcd}), $\frac{d \La}{dq} > 0$. Hence, all branches
$q_\ell (\La)$ of solutions of (\ref{laofq}) are monotonically
increasing, $\frac{d q_\ell (\La)}{d \La} > 0$. We can summarize the
properties of the solutions of equation (\ref{laofq}) in the following
lemma.
\begin{lemma}
If $L > \frac{8}{U}$, equation (\ref{laofq}) has exactly $L$ branches
of solutions $q_\ell (\La)$, which have the properties
\beq \begin{array}{r@{\qqd}l}
(i) & \dst{(\ell - 1) \frac{2\p}{L} \le q_\ell (\La) \le \ell
\frac{2\p}{L}} \qd, \\[2ex]
(ii) & \dst{\frac{d q_\ell (\La)}{d \La} > 0} \qd, \\[2ex]
(iii) & \dst{\lim_{\La \rightarrow - \infty} q_\ell (\La) =
(\ell - 1) \frac{2\p}{L} \qd, \qd
\lim_{\La \rightarrow \infty} q_\ell (\La) =
\ell \frac{2\p}{L}} \qd.
\end{array}
\eeq
\end{lemma}
(iii) can be seen from figure 1.
\begin{figure}
\begin{center}
\fbox{
\epsfxsize 8cm
\epsffile{fig1.eps}
}
\caption{Sketch of equation (\ref{laofq}).}
\end{center}
\end{figure}
Lemma III.1 is sufficient to characterize and count the real
solutions of (\ref{bak1}) and (\ref{bas1}) (or equivalently
(\ref{momlog}) and (\ref{laofk})). We are seeking for solutions of
(\ref{bak1}) and (\ref{bas1}) modulo permutations, where all $k_j$ are
mutually distinct. Choose $N$ branches $q_{\ell_1} (\La) < \dots
< q_{\ell_N} (\La)$ of solutions of (\ref{laofq}), and define
\beq
Q(\La) = q_{\ell_1} (\La) + \dots + q_{\ell_N} (\La) \qd.
\eeq
Then $k_j = q_{\ell_j} (\La)$, $j = 1, \dots, N$, and $\La$ solve
(\ref{bak1}) and (\ref{bas1}), if and only if
\beq \label{count}
Q(\La) \mod \, 2 \p = m \frac{2\p}{L} \qd.
\eeq
Now, using Lemma III.1,
\beq
\lim_{\La \rightarrow - \infty} Q(\La) = (\ell_1 + \dots +
\ell_N - N)
\frac{2\p}{L} \qd, \qd
\lim_{\La \rightarrow \infty} Q(\La) = (\ell_1 + \dots + \ell_N)
\frac{2\p}{L} \qd.
\eeq
Furthermore, $\frac{d Q(\La)}{d \La} > 0$. Thus there are precisely
$N - 1$ values $\La_\a$, $\a = 1, \dots, N - 1$, which satisfy
(\ref{count}) for a given choice $\ell_1 < \dots < \ell_N$ of branches
of equation (\ref{laofq}) (recall that we exclude $\La = \pm \infty$).
We summarize our result in the following
\begin{lemma}
Let $L > \frac{8}{U}$. Then there are precisely ${L \choose N} (N - 1)$
solutions with all $k_j$ real to equations (\ref{bak1}), (\ref{bas1}).
To every choice of $N$ mutually distinct intervals ${\cal I}_j =
[(\ell_j - 1) \frac{2 \p}{L}, \ell_j \frac{2 \p}{L}]$, $j = 1, \dots,
N$, $\ell_j \ne \ell_k$, there correspond $N - 1$ solutions with
$k_j \in {\cal I}_j$. These solutions are characterized by $N - 1$
different values of $\La$.
\end{lemma}
Let us show that our result coincides with Takahashi's counting
(\ref{r1})-(\ref{r3}). We have $N$ real $k$'s and one $\La$ string of
length 1. Since there is no $k$-$\La$ string, there is no ${J'}_\a^n$
to specify. $M' = 0$, $M_1 = 1$, $M_j = 0$ for $j > 1$. Thus $|J^1| \le
\2 (N - t_{11} - 1) = \2 (N - 2)$, and there are $N - 1$ possible
values of $J^1$. The number of different sets $\{I_1, \dots, I_N\}$
follows from (\ref{r1}) as $L \choose N$. This means that Takahashi's
counting predicts a total number of ${L \choose N} (N - 1)$ real
solutions, which is in accordance with Lemma III.2 as long as the
Takahashi condition (\ref{tcd}) is satisfied. In the special cases
$N = 2$ and $N = 3$ we find $L \choose 2$ and $2 {L \choose 3}$
solutions, respectively.
\subsection{\boldmath $k$-$\La$ two string}
Let us consider equation (\ref{laofk}) in the case that two of the
$k_j$'s are complex conjugated and the others are real. We may set
$k_- = k_1 = q - \i \x$, $k_+ = k_2 = q + \i \x$ with real $q$ and
real, positive $\x$. It follows from (\ref{laofk}) that
\beq
\sin(q + \i \x) = \sin(q) \ch(\x) + \i \cos(q) \sh(\x)
= \La + \frac{U}{4} \frac{\sin(qL) - \i \, \sh(\x L)}
{\ch(\x L) - \cos(qL)} \qd,
\eeq
or, if we separate real and imaginary part of this equation,
\bea \label{resin}
\sin(q) \ch(\x) & = & \La + \frac{U}{4}
\frac{\sin(qL)}{\ch(\x L) - \cos(qL)} \qd,
\\ \label{imsin}
\cos(q) \sh(\x) & = & - \frac{U}{4}
\frac{\sh(\x L)}{\ch(\x L) - \cos(qL)} \qd.
\eea
Note that for $\xi = 0$ equation (\ref{imsin}) is satisfied identically
and (\ref{resin}) turns into (\ref{laofq}).
Let us consider the two-electron case $N = 2$ first. Then, by equation
(\ref{momlog}),
\beq \label{momqn}
q = m \frac{\p}{L} \qd, \qd m = 0, \dots, 2L - 1 \qd.
\eeq
Equation (\ref{imsin}) determines $\x$ as a function of $q = m
\frac{\p}{L}$, and equation (\ref{resin}) determines $\La$. We have
$qL = m \p$. Hence, $\sin(qL) = 0$, $\cos(qL) = (- 1)^{m}$, and
(\ref{resin}) and (\ref{imsin}) decouple into
\bea \label{laq}
\La & = & \sin(q) \ch(\x) \\ \label{xiq}
\sh(\x) & = & - \, \frac{U}{4 \cos(q)} \,
\frac{\sh(\x L)}{\ch(\x L) - (- 1)^m} \qd.
\eea
This decoupling is a peculiarity of the two particle case, which makes
it more simple than the general case. We can discuss equation
(\ref{xiq}) graphically (see figure 2).
\begin{figure}
\begin{center}
\fbox{
\epsfxsize 8cm
\epsffile{fig2.eps}
}
\caption{Sketch of equation (\ref{xiq}).}
\end{center}
\end{figure}
Let us define
\beq \label{fofxi}
f(\xi) = \frac{\sh(\xi L)}{\ch(\xi L) - (- 1)^m} = \left\{
\begin{array}{l@{\qd, \qd \mbox{for $m$ }}l}
\tanh \left( \frac{\xi L}{2} \right) & \mbox{odd} \\[2ex]
\cth \left( \frac{\xi L}{2} \right) & \mbox{even} \qd.
\end{array} \right.
\eeq
$f(\xi) > 0$ for all $\xi > 0$. Hence, (\ref{xiq}) can have solutions
for positive $\xi$ only if
\beq \label{rangem}
\frac{\p}{2} < q < \frac{3 \p}{2} \qd, \qd \mbox{for\ } U > 0
\qd, \qd \frac{\p}{2} < |q - \p| < \p \qd, \qd \mbox{for\ }
U < 0 \qd.
\eeq
For the sake of simplicity let us concentrate on the repulsive case
$U > 0$. It follows from (\ref{fofxi}) that (\ref{xiq}) always has
exactly one solution $\xi_m$ for every even $m$ which satisfies
(\ref{rangem}). The condition for a solution with odd $m$ to exist is
that the derivative of the right hand side of equation (\ref{xiq}) is
larger than the derivative of the left hand side of equation (\ref{xiq})
as $\xi$ approaches zero from the right, i.e.
\beq
- \, \frac{UL}{8 \cos(q)} > 1 \qd.
\eeq
This is satisfied for all $q$ with $\frac{\p}{2} < q < \frac{3 \p}{2}$,
if and only if
\beq
L > \frac{8}{U} \qd.
\eeq
Again we have found the Takahashi condition (\ref{tcd}). If the
Takahashi condition is satisfied, then there is one and only one
$k$-$\La$ two string solution for every $m$ satisfying (\ref{momqn})
and (\ref{rangem}), and we can easily count these solutions. Their
number as a function of $L$ is different for odd and even $L$,
respectively.
\begin{lemma}
Let $L > \frac{8}{U}$, $U > 0$, $N = 2$. Equations (\ref{bak1}) and
(\ref{bas1}) have $k$-$\La$ two-string solutions only if the momentum
$q$ of the center of the string is in the range $\frac{\p}{2} < q
< \frac{3 \p}{2}$. The allowed values of $q$ in that range are
quantized as $q = m \frac{\p}{L}$. For every $q = m \frac{\p}{L}$
there is one and only one $k$-$\La$ two string. The total number of
$k$-$\La$ two strings is $L$ for $L$ odd and $L - 1$ for $L$ even.
\end{lemma}
Comparing our result with the prediction of Takahashi's counting
(\ref{r1})-(\ref{r3}) we find again agreement. Now there is no free
$k_j$ and no $\La$ string, thus no $I_j$ and no $J_\a^n$. Furthermore,
$M_1' = M' = 1$, and $M_j = 0$ for $j > 1$. Thus $|{J'}^1| \le
\2 (L - t_{11} - 1) = \2 (L - 2)$, which means that there are $L - 1$
possible values of ${J'}^1$. This agrees with our Lemma since $L$ was
assumed to be even in (\ref{r1})-(\ref{r3}).
Note that $\lim_{L \rightarrow \infty} f(\xi) = 1$, pointwise for all
$\xi > 0$. This suggests the notion of an ideal string determined by
the equations
\beq \label{ideal2}
\La = \sin(q) \ch(\x) \qd, \qd
\sh(\x) = - \, \frac{U}{4 \cos(q)} \qd, \qd
q = m \frac{\p}{L} \qd.
\eeq
Replacing (\ref{laq}) and (\ref{xiq}) by (\ref{ideal2}) means to
replace the curves denoted by `$m$ odd' and `$m$ even' in figure 2
by the dashed line. Clearly, solutions of (\ref{laq}), (\ref{xiq})
are in one-to-one correspondence to solutions of (\ref{ideal2}), if
the Takahashi condition is satisfied. We can formulate the following
corrolary of Lemma III.3.
\begin{lemma}
Let $L > \frac{8}{U}$, $N = 2$. Let $m_L$ be a sequence of integers
($\frac{L}{2} < m_L < \frac{3L}{2}$), such that $\lim_{L \rightarrow
\infty} m_L \frac{\p}{L} = q$, $\frac{\p}{2} \le q \le \frac{3 \p}{2}$.
According to Lemma III.3 this defines a sequence $\xi_{m_L}$ of
solutions of (\ref{xiq}). This sequence has the limit
\beq \label{limes2}
\lim_{L \rightarrow \infty} \xi_{m_L} =
- \, \arsh \left( \frac{U}{4 \cos(q)} \right) \qd,
\eeq
i.e.\ in the thermodynamic limit all $k$-$\La$ two strings are driven
to the ideal string positions.
\end{lemma}
Proof: (\ref{limes2}) follows from (\ref{xiq}), since the sequence
$\xi_{m_L}$ is bounded from below. Let us prove the latter statement.
Assume the contrary. Then there is a subsequence $\xi_{m_{L_j}}$
of $\xi_{m_L}$ which goes to zero, and it follows from (\ref{xiq})
that $\lim_{j \rightarrow \infty} f(\xi_{n_{L_j}}) = 0$. This means
(i) $m_{L_j}$ is odd for all sufficiently large $j$, and (ii)
$\lim_{j \rightarrow \infty} \sh(\xi_{m_{L_j}} L_j) =
\lim_{j \rightarrow \infty} \xi_{m_{L_j}} L_j = 0$. We conclude that
\beq
\lim_{j \rightarrow \infty}
\frac{\sh(\xi_{m_{L_j}})}{\sh(\xi_{m_{L_j}} L_j)} =
\lim_{j \rightarrow \infty}
\frac{1}{L_j} = 0 =
\lim_{j \rightarrow \infty}
- \frac{U}{8 \cos(m_{L_j} \frac{2 \p}{L_j})} =
- \frac{U}{8 \cos(q)} \qd,
\eeq
which is a contradiction. Thus, the lemma is proved.
Let us now proceed with the case $N = 3$. We have to solve the following
system of equations (cf.\ (\ref{momlog}), (\ref{laofk}), (\ref{resin}),
(\ref{imsin})),
\bea \label{n31}
2 q + k_3 & = & m \frac{2 \p}{L} \qd, \qd m = 0, \dots, 3L -1
\qd, \\ \label{n32}
\sin(k_3) - \La & = & \frac{U}{4}
\ctg \left( \frac{k_3 L}{2} \right)
\qd, \\ \label{n33}
\La & = & \sin(q) \ch(\xi) - \frac{U}{4} \frac{\sin(qL)}
{\ch(\xi L) - \cos(qL)} \qd, \\ \label{n34}
\cos(q) \sh(\xi) & = & - \frac{U}{4} \frac{\sh(\xi L)}
{\ch(\xi L) - \cos(qL)} \qd.
\eea
We choose a branch of solution $q_\ell (\La)$ of (\ref{n32}) and insert
it into (\ref{n31}). This yields
\beq \label{cmm}
q = m \frac{\p}{L} - \frac{q_\ell (\La)}{2} \qd.
\eeq
Using this result, (\ref{n33}) and (\ref{n34}) turn into
\bea \label{la3}
\La & = & \sin \left(\tst{m \frac{\p}{L} - \frac{q_\ell (\La)}{2}}
\right) \ch(\xi) - \frac{U}{4}
\frac{\sin(q_\ell (\La) L/2)}
{\cos(q_\ell (\La) L/2) - (- 1)^m \ch(\xi L)}
\qd, \\ \label{xi3}
\cos \left(\tst{m \frac{\p}{L} - \frac{q_\ell (\La)}{2}} \right)
\sh(\xi) & = & - \frac{U}{4} \frac{\sh(\xi L)}
{\ch(\xi L) - (- 1)^m \cos(q_\ell (\La) L/2)}
\qd.
\eea
This is a system of two equations in two unknowns $\xi$ and $\La$. In
contrast to the case $N = 2$, which was considered above, these
equations do not decouple.
We shall first consider the solutions $\xi_{\ell, m} (\La)$ of equation
(\ref{xi3}). For this purpose we define
\beq
f_a (\xi) = \frac{\sh(\xi L)}{\ch(\xi L) - a} \qd, \qd
a = (- 1)^m \cos \left( \frac{q_\ell (\La) L}{2} \right) \qd, \qd
b = - \, \frac{U}{4 \cos \left( \tst{m \frac{\p}{L} -
\frac{q_\ell (\La)}{2}} \right)} \qd.
\eeq
With these definitions equation (\ref{xi3}) turns into
\beq \label{xi3ofla}
\sh(\xi) = b f_a (\xi) \qd,
\eeq
Note that $a$ is real and $|a| \le 1$. Hence, $f_a (\xi) > 0$ for all
positive $\xi$, and a necessary condition for (\ref{xi3ofla}) to have
a solution is $b > 0$. This means that
\beq \label{rangeq}
\frac{\p}{2} \le m \frac{\p}{L} - \frac{q_\ell (\La)}{2} \le
\frac{3 \p}{2} \qd, \qd \mbox{for\ } U > 0
\qd, \qd \frac{\p}{2} \le \left| m \frac{\p}{L} -
\frac{q_\ell (\La)}{2} - \p \right| \le \p
\qd, \qd \mbox{for\ } U < 0 \qd.
\eeq
Let us concentrate again on the case $U > 0$. The inequality for the
case $U > 0$ in (\ref{rangeq}) holds for all real $\La$, if and only
if $\frac{L}{2} \le m - \ell \le \frac{3L}{2} - 1$ (cf.\ Lemma III.1).
Equation (\ref{xi3ofla}) can be easily solved graphically for fixed $a$
($|a| < 1$) and $b > 0$ (see figure 3). The function $f_a (\xi)$ has
the following properties. (i) $a < 0$: $f_a (\xi)$ is monotonically
increasing ($f_a' (\xi) > 0$) and concave ($f_a'' (\xi) < 0$),
$f_a (0) = 0$, $f_a' (0) = \frac{L}{1 - a} \ge \frac{L}{2}$ and
$\lim_{\xi \rightarrow \infty} f_a (\xi) = 1$. (ii) $a > 0$: $f_a (\xi)$
has a single positive maximum $\xi^{(0)}$ and a single positive
turning point $\xi^{(1)}$, $\xi^{(0)} < \xi^{(1)}$, $f_a (0) = 0$,
$f_a' (0) = \frac{L}{1 - a} \ge \frac{L}{2}$ and $\lim_{\xi \rightarrow
\infty} f_a (\xi) = 1$. These properties are sufficient to conclude
that (\ref{xi3ofla}) has a unique solution $\xi_{\ell, m} (\La)$ for
all real $\La$, if and only if $\frac{L}{2} \le m - \ell \le
\frac{3L}{2} - 1$ (recall that $U > 0$) and the Takahashi condition
$L > \frac{8}{U}$ is satisfied. Note that $f_a (\xi)$ as a function of
$a$ interpolates between the two branches of the function $f(\xi)$,
equation (\ref{fofxi}). These two branches are the two dashed lines
envelopping the functions $f_a (\xi)$ in figure 3.
\begin{figure}
\begin{center}
\fbox{
\epsfxsize 8cm
\epsffile{fig3.eps}
}
\caption{Sketch of equation (\ref{xi3ofla}).}
\end{center}
\end{figure}
Lemma III.1 and Lemma III.3 allow us to understand the behaviour of the
solutions $\xi_{\ell, m} (\La)$ of (\ref{xi3}) as $\La \rightarrow
\pm \infty$. Lemma III.1 applied to (\ref{xi3}) implies
\bea
\cos \left((m - \ell + 1) \tst{\frac{\p}{L}} \right) \sh(\xi) & = &
- \frac{U}{4} \frac{\sh(\xi L)}{\ch(\xi L) - (- 1)^{m - \ell + 1}}
\qd, \qd \mbox{for\ } \La \rightarrow - \infty \qd, \\
\cos \left((m - \ell) \tst{\frac{\p}{L}} \right) \sh(\xi) & = &
- \frac{U}{4} \frac{\sh(\xi L)}{\ch(\xi L) - (- 1)^{m - \ell}}
\qd, \qd \mbox{for\ } \La \rightarrow \infty \qd.
\eea
Using Lemma III. 3 we conclude that
\beq \label{limxila}
\lim_{\La \rightarrow - \infty} \xi_{\ell, m} (\La) =
\xi_{m - \ell + 1} \qd, \qd
\lim_{\La \rightarrow \infty} \xi_{\ell, m} (\La) =
\xi_{m - \ell} \qd,
\eeq
where $\xi_m$ is the unique solution of equation (\ref{xiq}). This
solution exists (cf.\ (\ref{rangem}) and recall that $U > 0$), if and
only if $\frac{L}{2} < m < \frac{3L}{2}$. Hence, the range of validity
of (\ref{limxila}) is restricted to
\beq \label{rangexila}
\frac{L}{2} < m - \ell < \frac{3L}{2} - 1 \qd.
\eeq
Let us insert $\xi_{\ell, m} (\La)$ into (\ref{la3}). Then (\ref{la3})
turns into
\beq \label{lag}
\La = g_{\ell, m} (\La) \qd,
\eeq
where $g_{\ell, m} (\La)$ is defined by
\beq \label{defg}
g_{\ell, m} (\La) = \sin \left(\tst{m \frac{\p}{L} -
\frac{q_\ell (\La)}{2}} \right)
\ch(\xi_{\ell, m} (\La)) -
\frac{U}{4} \frac{\sin(q_\ell (\La) L/2)}
{\cos(q_\ell (\La) L/2) -
(- 1)^m \ch(\xi_{\ell, m} (\La) L)} \qd.
\eeq
Using (\ref{limxila}) we obtain the asymptotics of $g_{\ell, m} (\La)$,
\beq \label{limgla}
\lim_{\La \rightarrow - \infty} g_{\ell, m} (\La) =
\sin \left( (m - \ell + 1) \tst{\frac{\p}{L}} \right)
\ch(\xi_{m - \ell + 1}) \qd, \qd
\lim_{\La \rightarrow \infty} g_{\ell, m} (\La) =
\sin \left( (m - \ell) \tst{\frac{\p}{L}} \right)
\ch(\xi_{m - \ell}) \qd.
\eeq
We see that $g_{\ell, m} (\La)$ has finite asymptotics for $\La
\rightarrow \pm \infty$. Since $g_{\ell, m} (\La)$ is continuous in
$\La$, we arrive at the conclusion that there exists a solution
$\La_{\ell, m}$ of (\ref{lag}) for every pair $(\ell, m)$, which
satisfies (\ref{rangexila}). Hence, we have shown the following
\begin{lemma}
Let $L > \frac{8}{U}$, $U > 0$, $N = 3$. Equations (\ref{bak1}) and
(\ref{bas1}) have solutions consisting of one $k$-$\Lambda$ two string
and a single real $k$ only if the momentum $q$ of the center of the
string is in the range $\frac{\p}{2} < q < \frac{3 \p}{2}$. For every
choice of branch of the real momentum $k_3$ (cf.\ figure 1), there
exist $L - 1$ $k$-$\La$ two strings for odd $L$ and $L - 2$ $k$-$\La$
two strings for even $L$. This gives a total number of $L(L - 1)$
solutions of this type for $L$ odd and of $L(L - 2)$ for $L$ even,
respectively.
\end{lemma}
Let us compare with Takahashi's counting (\ref{r1})-(\ref{r3}). We
have a single free $k_j$ and no $\La$ string, which means that there
is one $I_j = I$ and no $J_\a^n$. It follows from (\ref{r1}) that
$I$ may take $L$ different values. Furthermore, $M_1' = M' = 1$, and
$M_j = 0$ for $j > 1$. Thus $|{J'}^1| \le \2 (L - 3)$, which leads to
$L - 2$ possible values of ${J'}^1$. Takahashi's counting therefore
gives $L(L - 2)$ solutions with one $k$-$\Lambda$ string and one real
$k$. This is in aggreement with our Lemma.
We are now ready to state the following generalization of Lemma III.4,
\begin{lemma}
Let $L > \frac{8}{U}$, $N = 3$. Choose two sequences of integers
$\ell_L$ and $m_L$, such that $\lim_{L \rightarrow \infty}
(m_L - \ell_L) \frac{\p}{L} = q$, $\frac{\p}{2} \le q \le
\frac{3 \p}{2}$. This defines a sequence of solutions
$\xi_{\ell_L, m_L} (\La)$ of equation (\ref{xi3}), which has the limit
\beq
\lim_{L \rightarrow \infty} \xi_{\ell_L, m_L} (\La) =
- \, \arsh \left( \frac{U}{4 \cos(q)} \right) \qd,
\eeq
uniformly in $\La$. Thus all strings corresponding to the sequence
$\xi_{\ell_L, m_L} (\La)$ are driven to their ideal positions.
\end{lemma}
Proof: There is no $\La$ and no subsequence $\xi_{\ell_{L_j},
m_{L_j}} (\La)$, such that $\lim_{j \rightarrow \infty}
\xi_{\ell_{L_j}, m_{L_j}} (\La) = 0$. This can be seen in similar way
as in the proof of Lemma III.4.
Let us finally note that our considerations for the case $N = 3$
readily generalize to arbitrary $N$. For arbitrary $N$ we have to
consider $N - 2$ copies of equation (\ref{n32}) in the system of
equations (\ref{n31})-(\ref{n34}). We further have to replace
$q_\ell (\La)$ in (\ref{cmm}) by $Q(\La) = q_{\ell_1} (\La) + \dots
+ q_{\ell_{N - 2}} (\La)$ with $\ell_1 < \dots < \ell_{N - 2}$. The
properties of $Q(\La)$ (monotonicity and asymptotics) then follow from
Lemma III.1, and all considerations go through as in the case $N = 3$.
\subsection{Summary}
In this section we have studied $k$-$\La$ string solutions of the
Lieb-Wu equations (\ref{bak}), (\ref{bas}). We have shown that such
solutions exist and that they are driven to certain ideal string
positions in the limit of a large lattice. We have further shown that
for a large enough lattice of finite length their number is in
accordance with the number of corresponding solutions of the discrete
Takahashi equations (\ref{t1})-(\ref{t3}).
\section{Lieb-Wu equations for a single down spin (II) --
Self-consistent solution}
\subsection{Zeroth order and the discrete Takahashi equations}
In this section we present the self-consistent solution of the Lieb-Wu
equations (\ref{bak}), (\ref{bas}) for the case of three electrons
and one $k$-$\La$ string. The Lieb-Wu equations (\ref{bak}), (\ref{bas})
provide a self-consistent way of calculation of the deviation of the
strings from their ideal positions. We show that every solution of the
discrete Takahashi equations gives an approximate solution to the
Lieb-Wu equations (\ref{bak}), (\ref{bas}), and we calculate the leading
order corrections. These corrections vanish exponentially fast as the
number of lattice sites $L$ becomes large.
The Lieb-Wu equations for three electrons and one down spin are
\bea \label{bak3}
&& e^{\i k_j L} = \frac{\La - \sin k_j - \i U/4}
{\La - \sin k_j + \i U/4} \qd, \qd
j = 1, 2, 3 \qd,\\ \label{bas3}
&& \prod_{j = 1}^3 \frac{\La - \sin k_j - \i U/4}
{\La - \sin k_j + \i U/4} = 1 \qd.
\eea
As in section III.A we may replace equation (\ref{bas3}) by the
equation for the conservation of momentum,
\beq \label{emom3}
e^{\i (k_1 + k_2 + k_3)L} = 1 \qd.
\eeq
(\ref{bak3}) and (\ref{emom3}) follow from (\ref{bak3}) and (\ref{bas3})
and vice versa.
Let us follow the usual self-consistent strategy for obtaining a
$k$-$\La$ string solution. As in section III.C we shall use the
notation $k_- = k_1 = q - \i \xi$, $k_+ = q + \i \xi$, $\xi > 0$,
i.e.\ we assume that $k_1$ and $k_2$ are part of a $k$-$\La$ string.
In order to facilitate comparison with the previous literature
we shall also use the abbreviations $\Ph = \Re \sin(k_+)
= \Re \sin(k_-)$ and $\chi = \Im \sin(k_-) = - \Im \sin(k_+)$. We
introduce $\de$ as a measure of the deviation of the string from its
ideal position. Then
\beq \label{dev}
\sin(k_-) = \Ph + \i \chi = \La + \frac{\i U}{4} + \de \qd, \qd
\sin(k_+) = \Ph - \i \chi = \La - \frac{\i U}{4} + \bar \de \qd.
\eeq
Inserting (\ref{dev}) into (\ref{bak3}) gives
\beq \label{epm}
e^{\i k_- L} = 1 + \frac{\i U}{2 \de} \qd, \qd
e^{- \i k_+ L} = 1 - \frac{\i U}{2 \bar \de} \qd.
\eeq
We may consider the first equation in (\ref{dev}) as the equation, that
defines $\de$. The second equation in (\ref{dev}) is not independent.
It is the complex conjugated of the first one and may be dropped for
that reason. Similarly, we may also drop the second equation in
(\ref{epm}). Then we are left with six independent equations,
(\ref{bak3}) for $j = 3$, (\ref{emom3}), and the real and imaginary
parts of the first equations in (\ref{dev}) and (\ref{epm}). Note that
$k_1 + k_2 = 2q$. Therefore our six equations are equivalent to
\bea \label{basc1}
e^{2 \i q L} & = & \frac{\La - \sin k_3 + \i U/4}
{\La - \sin k_3 - \i U/4} \qd, \\ \label{basc2}
e^{\i k_3 L} & = & \frac{\La - \sin k_3 - \i U/4}
{\La - \sin k_3 + \i U/4} \qd, \\ \label{basc3}
\sin(k_-) & = & \La + \frac{\i U}{4} + \de \qd, \\ \label{basc4}
\de & = & \frac{\i U}{2} \, \frac{1}{e^{\i k_- L} - 1} \qd.
\eea
Every $k$-$\La$ string solution of the Lieb-Wu equations (\ref{bak3}),
(\ref{bas3}) gives a solution of equations (\ref{basc1})-(\ref{basc4})
(with real $q$, $k_3$, $\La$ and real positive $\xi$) and vice versa.
If $L$ is large and $\xi = \Im k_+ = \CO (1)$, then $\de$ is very
small and may be neglected in equation (\ref{basc3}). Then (\ref{basc4})
decouples from the other equations, which become
\bea \label{pt1}
e^{2 \i q^{(0)} L} & = & \frac{\La_0 - \sin k_3^{(0)} + \i U/4}
{\La_0 - \sin k_3^{(0)} - \i U/4} \qd, \\
\label{pt2}
e^{\i k_3^{(0)} L} & = & \frac{\La_0 - \sin k_3^{(0)} - \i U/4}
{\La_0 - \sin k_3^{(0)} + \i U/4} \qd, \\
\label{pt3}
\sin(k_-^{(0)}) & = & \La_0 + \frac{\i U}{4} \qd.
\eea
If, on the other hand, $\de$ is very small in (\ref{basc3}), we may
neglect it in first approximation and solve (\ref{pt3}) instead. Now
(\ref{pt3}) implies that $\xi^{(0)} = \CO (1)$ (see below). Then, using
(\ref{basc4}), we see that $\de$ is indeed small for large $L$. This
means the assumption $\de$ be small for large $L$ is self-consistent.
Let us be more precise. We set $k_-^{(0)} = q^{(0)} - \i \xi^{(0)}$ with
real $q^{(0)}$ and real, positive $\xi^{(0)}$. Separating (\ref{pt3})
into real and imaginary part we obtain two equations which relate the
three unknowns $q^{(0)}$, $\xi^{(0)}$ and $\La_0$,
\bea \label{idla}
\La_0 & = & \sin(q^{(0)}) \ch(\xi^{(0)}) \qd, \\ \label{idxi}
\xi^{(0)} & = & - \, \arsh \left( \frac{U}{4 \cos(q^{(0)})} \right)
\qd.
\eea
Let us concentrate on positive coupling $U > 0$ for simplicity. Since
we are assuming that $\xi^{(0)} > 0$, the range of $q^{(0)}$ is then
restricted to, say, $\frac{\p}{2} < q^{(0)} < \frac{3 \p}{2}$ by
equation (\ref{idxi}). We obtain the uniform estimate
\beq \label{xiest}
\xi^{(0)} > \arsh(U/4) \qd.
\eeq
In order to test, if a solution $q^{(0)}$, $\xi^{(0)}$, $k_3^{(0)}$,
$\La_0$ of (\ref{pt1})-(\ref{pt3}) is a good approximate solution of
(\ref{basc1})-(\ref{basc4}), we use it to estimate the modulus of
$\de$ for large $L$,
\beq \label{deest}
|\de| \approx \frac{U}{2} \, e^{- \xi^{(0)} L} <
\frac{U}{2} \, e^{- \arsh(U/4) L} \qd.
\eeq
The inequality follows from (\ref{xiest}). We conclude that $|\de|$
becomes very small for large $L$. For $U = 4$ and $L = 24$, for
instance, the estimate (\ref{deest}) gives $|\de| < 1.3 \cdot 10^{- 9}$,
whereas the difference between two real $k_3$'s is of the order of
$\frac{2 \p}{L} = 0.26$ (cf.\ section III.B). For large $L$ it becomes
impossible to numerically distinguish between solutions of
(\ref{basc1})-(\ref{basc4}) and (\ref{pt1})-(\ref{pt3}), respectively.
If we fix the branch of the arcsin as $- \frac{\p}{2} \le \arcsin (z)
\le \frac{\p}{2}$, it follows from the inequality $\frac{\p}{2} <
q^{(0)} < \frac{3 \p}{2}$ that
\beq \label{kmla}
k_-^{(0)} = \p - \arcsin \left( \La_0 + \tst{\frac{\i U}{4}} \right)
\qd.
\eeq
Inserting (\ref{kmla}) into (\ref{pt1}) leads to
\beq \label{ptp}
e^{2 \i \, \Re \arcsin(\La_0 + \i U/4) L} =
\frac{\La_0 - \sin k_3^{(0)} - \i U/4}
{\La_0 - \sin k_3^{(0)} + \i U/4}
\qd.
\eeq
We thus have eliminated $k_-^{(0)}$ from the system of equations
(\ref{pt1})-(\ref{pt3}), and we are left with two independent equations
(\ref{pt2}) and (\ref{ptp}). These two equations determine the two
real unknowns $\La_0$ and $k_3^{(0)}$. Taking logarithms of (\ref{pt2})
and (\ref{ptp}) we arrive at Takahashi's discrete equations (\ref{t1}),
(\ref{t3}) for one $k$-$\La$ string and one real $k$.
We have seen that the equations (\ref{pt1})-(\ref{pt3}) determine
Takahashi's {\it ideal strings}. Equations (\ref{basc1})-(\ref{basc2})
on the other hand, are equations for {\it non-ideal strings}, which
solve the Lieb-Wu equations (\ref{bak3}), (\ref{bas3}). Thus, $\de$ is
a measure for the deviation of the strings from their ideal positions.
We have further seen that the assumption that $\de$ be small is
self-consistent. In particular, every solution of the equations
(\ref{pt1})-(\ref{pt3}), which are equivalent to Takahashi's discrete
equations, is an approximate solution of equations (\ref{basc1})-%
(\ref{basc4}). The approximation becomes extremely accurate for large
$L$.
\subsection{First order corrections}
Inserting any solution of (\ref{pt1})-(\ref{pt3}) into (\ref{basc4})
we have
\beq \label{deexp}
\de = \frac{\i U}{2} \, e^{- \i q^{(0)} L} e^{- \xi^{(0)} L} +
\CO \left(e^{- 2 \xi^{(0)} L} \right) =
\frac{U}{2} \, \sin(q^{(0)} L) e^{- \xi^{(0)} L} +
\frac{\i U}{2} \, \cos(q^{(0)} L) e^{- \xi^{(0)} L} +
\CO \left(e^{- 2 \xi^{(0)} L} \right) \qd.
\eeq
So the relevant parameter, which controls the deviation of the strings
from their ideal positions is $\epsilon = e^{- \xi^{(0)} L}$. Every
solution of (\ref{pt1})-(\ref{pt3}) is an approximate solution of
(\ref{basc1})-(\ref{basc4}). Let us calculate the leading order
corrections. We expect them to be proportional to $\epsilon$,
\bea \label{qeps}
q & = & q^{(0)} + q^{(1)} \epsilon + \CO (\epsilon^2) \qd, \\ \label{xieps}
\xi & = & \xi^{(0)} + \xi^{(1)} \epsilon + \CO (\epsilon^2) \qd, \\
\label{k3eps}
k_3 & = & k_3^{(0)} + k_3^{(1)} \epsilon + \CO (\epsilon^2) \qd, \\
\label{laeps}
\La & = & \La_0 + \La_1 \epsilon + \CO (\epsilon^2) \qd.
\eea
The $\epsilon$-expansion of $\de$ is given in equation (\ref{deexp}) above.
We also introduce
\bea \label{phieps}
\Ph & = & \sin(q^{(0)}) \ch(\xi^{(0)}) + \Ph^{(1)} \epsilon
+ \CO (\epsilon^2) \qd, \\
\chi& = & \frac{U}{4} + \chi^{(1)} \epsilon + \CO (\epsilon^2) \qd,
\label{chieps}
\eea
since it is the quantities $\Ph = \Re \sin(k_-)$ and $\chi =
\Im \sin(k_-)$ which actually form strings in the complex plane.
The two sets of variables $q$, $\xi$ and $\Ph$, $\chi$ are not
independent. Inserting (\ref{qeps}) and (\ref{xieps}) into the left
hand side of (\ref{phieps}) and (\ref{chieps}) and comparing leading
orders in $\epsilon$ we find
\beq \label{li1}
\left( \begin{array}{c} \Ph^{(1)} \\ \chi^{(1)} \end{array}
\right) =
\left( \begin{array}{cc} \cos(q^{(0)}) \ch(\xi^{(0)}) &
\sin(q^{(0)}) \sh(\xi^{(0)})\\ \sin(q^{(0)}) \sh(\xi^{(0)}) &
- \cos(q^{(0)}) \ch(\xi^{(0)}) \end{array} \right)
\left( \begin{array}{c} q^{(1)} \\ \xi^{(1)} \end{array} \right)
\qd.
\eeq
Let us insert (\ref{deexp}), (\ref{phieps}) and (\ref{chieps}) into
(\ref{basc3}). We obtain to leading order
\bea \label{li2}
\Ph^{(1)} & = & \La_1 + \frac{U}{2} \sin(q^{(0)} L) \qd, \\
\chi^{(1)} & = & \frac{U}{2} \cos(q^{(0)} L) \qd, \label{li3}
\eea
i.e.\ we have already found the leading order correction $\chi^{(1)}$,
equation (\ref{li3}). Next we insert (\ref{qeps}), (\ref{k3eps}) and
(\ref{laeps}) into (\ref{basc1}) and (\ref{basc2}) and linearize in
$\epsilon$. The resulting equations are
\bea \label{li4}
2 q^{(1)} & = & - \frac{U}{2L} \,
\frac{\La_1 - \cos(k_3^{(0)}) \, k_3^{(1)}}
{\left(\La_0 - \sin(k_3^{(0)})\right)^2
+ \frac{U^2}{16}} \qd, \\ \label{li5}
k_3^{(1)} & = & \frac{U}{2L} \,
\frac{\La_1 - \cos(k_3^{(0)}) \, k_3^{(1)}}
{\left(\La_0 - \sin(k_3^{(0)})\right)^2 +
\frac{U^2}{16}} = - 2 q^{(1)} \qd.
\eea
The equation $2 q^{(1)} + k_3^{(1)} = 0$ following from (\ref{li4}),
(\ref{li5}) is, of course, a consequence of momentum conservation.
Equations (\ref{li1})-(\ref{li5}) are a system of six linear equations
for six unknowns, $q^{(1)}$, $\xi^{(1)}$, $\Ph^{(1)}$, $\chi^{(1)}$,
$k_3^{(1)}$ and $\La_1$. Note that in the derivation of these equations
we have used all of the equations (\ref{basc1})-(\ref{basc4}).
Equations (\ref{li2}) and (\ref{li3}) came out of (\ref{basc3}),
(\ref{basc4}) was used to obtain the expansion (\ref{deexp}) for $\de$
in terms of $\epsilon$, and (\ref{li4}), (\ref{li5}) follow from
(\ref{basc1}) and (\ref{basc2}). Equation (\ref{li1}) is a consequence
of the definition of $\Ph$ and $\chi$. Being a set of linear equations,
(\ref{li1})-(\ref{li5}) are readily solved,
\bea \label{pert1}
\La_1 & = & - \frac{U}{2} \left( \sin(q^{(0)} L) + \tan (q^{(0)})
\tanh(\xi^{(0)}) \cos(q^{(0)} L) \right) C(L) \\
\label{pert2}
& = & - \frac{U}{2} \left( \sin(q^{(0)} L) + \tan (q^{(0)})
\tanh(\xi^{(0)}) \cos(q^{(0)} L) \right)
+ \CO(1/L) \qd, \\ \label{pert3}
k_3^{(1)} & = & \frac{\La_1}{\frac{2L}{U} \left[ \left( \La_0
- \sin(k_3^{(0)}) \right)^2 + \frac{U^2}{16}
\right] + \cos(k_3^{(0)})} \\ \label{pert4}
& = & - \, \frac{U^2 \left( \sin(q^{(0)} L) + \tan (q^{(0)})
\tanh(\xi^{(0)}) \cos(q^{(0)} L) \right)}
{4L \left[ \left( \La_0 - \sin(k_3^{(0)})
\right)^2 + \frac{U^2}{16}
\right]} + \CO(1/L^2) \qd, \\ \label{pert5}
\Ph^{(1)} & = & \La_1 + \frac{U}{2} \sin(q^{(0)} L) =
- \frac{U}{2} \tan (q^{(0)}) \tanh(\xi^{(0)})
\cos(q^{(0)} L) + \CO(1/L) \qd, \\ \label{pert6}
\chi^{(1)} & = & \frac{U}{2} \cos(q^{(0)} L) \qd.
\eea
The function $C(L)$ in equation (\ref{pert1}) which gives the explicit
form of the $\CO(1/L)$ and $\CO(1/L^2)$ corrections in the remaining
equations is
\beq
C(L) = \left[ 1 + \frac{\cos(q^{(0)}) \ch(\xi^{(0)})
(1 + \tan^2 (q^{(0)}) \tanh^2 (\xi^{(0)}))}
{\frac{4L}{U} \left[ \left( \La_0
- \sin(k_3^{(0)}) \right)^2 + \frac{U^2}{16}
\right] + 2 \cos(k_3^{(0)})} \right]^{- 1}
= 1 + \CO(1/L) \qd.
\eeq
Equations (\ref{pert1})-(\ref{pert6}) give a complete description of
the leading order deviation of a non-ideal string from its ideal
position in the presence of one real $k$. The deviations of $q$ and
$\xi$, which are determined by $q^{(1)}$ and $\xi^{(1)}$, follow from
the equations (\ref{li1}) and (\ref{pert5}), (\ref{pert6}). In order to
see that $C(L) - 1$ is indeed of order $\CO (1/L)$ on has to use
(\ref{idla}) and (\ref{idxi}).
\subsection{Summary}
In this section we have presented a self-consistent solution of the
Lieb-Wu equations for the case of three electrons and one $k$-$\La$
string. Recall that the existence of these solutions was shown in the
previous section. Here we showed that a self-consistent approach
naturally leads to Takahashi's discrete equations. We showed that
Takahashi's discrete equations provide a highly accurate approximate
solution of the Lieb-Wu equations in the limit of a large lattice.
We also showed that there is a natural parameter $\epsilon =
e^{- \xi^{(0)} L}$ that measures the deviation of solutions of the
Lieb-Wu equations from the corresponding solutions of the discrete
Takahashi equations. Employing an algebraic perturbation theory we
explicitly calculated the leading order deviation in $\epsilon$ of a
non-ideal $k$-$\La$ string from its ideal position in the presence of
a real $k$.
\section{Lieb-Wu equations for a single down spin (III) -- Numerical
solution}
\subsection{Numerical method}
In section III the existence of $k$-$\La$ string solutions was
analytically proven under the Takahashi condition for the simplest
non-trivial cases with one down spin, $M=1$. The deviation of ideal
string solutions given by the discrete Takahashi equations from the
corresponding solutions of the Lieb-Wu equations was evaluated
analytically in section IV, and it was shown that the corrections
vanish exponentially fast as the lattice size $L$ becomes large.
To confirm these analytical arguments on the existence of $k$-$\La$
strings, we utilize a complementary numerical approach. Although
the tractable system size is limited, we can directly obtain $k$-$\La$
strings and verify the completeness of the Bethe ansatz for arbitrary
$U$. We note that numerical solutions for low-lying particle-hole
and $\Lambda$ string excitations in small finite systems can be found
in the literature (see e.g.\ \cite{Schulz93}). As far as we know,
however, our data are the first example of a numerical study that
confirms the completeness of the Bethe ansatz for a small finite system.
We study again the cases $N=2$ and $N=3$ for both pedagogical clarity
and technical simplicity. We shall employ two numerically exact
methods, (1) the numerical diagonalization of a real-symmetric matrix
using the Householder-QR method (we call it Method 1), (2) a numerical
method to solve coupled nonlinear equations using the Brent
method (we call it Method 2). These techniques themselves are
conventional. They allow us to obtain numerically exact solutions for
the Hubbard model. Here we use the term `numerically exact' to state
that our numerical solutions give exact numbers except for the
inevitable rounding error in the computation.%
\footnote{We will indicate errors by the difference of the left hand
side and the right side of each equation in (\ref{bak}), (\ref{bas}),
evaluated within our numerical treatment. Then it will become clear that
in all the equations which we tested the relative error is negligible
and of the order of the rounding error expected for the double-precision
calculation.}
Our strategy here is the following.
\begin{enumerate}
\item Obtain all energy eigenvalues with fixed $N$ and $M$ as a
function of $U$ using Method 1. This step gives a complete list of
energy eigenvalues.
\item Obtain numerical (real and/or complex) solutions by solving the
Lieb-Wu equations with Method 2.
\item List up all eigenvalues obtained by the two methods and compare
them with one another. This step gives a confirmation of completeness.
\end{enumerate}
For our numerical study we used the following form of the Hubbard
Hamiltonian,
\beq
H = - \sum_{j=1}^L \sum_{\sigma= \uparrow, \downarrow}
\left( c_{j+1, \sigma}^+ c_{j, \sigma}
+ c_{j-1, \sigma}^+ c_{j, \sigma} \right)
+ U \sum_{j=1}^{L} n_{j, \uparrow}
n_{j, \downarrow} \qd.
\label{hamiltonian}
\eeq
This form is different from (\ref{ham}) by a shift of the chemical
potential and by a constant energy shift. For fixed particle number
$N$ this leads to a shift of the spectrum by $\frac{U}{4} (2N - L)$.
In order to perform a numerical diagonalization we used basis vectors
$c_{x_1}^+ \dots c_{x_{N - M}}^+ c_{y_1}^+ \dots
c_{y_M}^+ |0\>$ to
represent the Hamiltonian as a matrix in the sector of fixed $N$ and
$M$. The number of different configurations, $x_1 < \dots < x_{N - M}$,
$y_1 < \dots < y_M$, in this sector is ${L \choose N - M} \times
{L \choose M}$ and determines the size of the matrix. In order to
employ method 2, we rewrote the Lieb-Wu equations into a proper
set of real equations, which will be presented in the following
subsections. Hereafter in this section we assume that $L$ is an even
integer.
\subsection{Numerical solution for two electrons}
\subsubsection{Equations for the $k$-$\Lambda$ string}
Let us discuss numerical $k$-$\La$ two string solutions for the two
electron system with one down spin ($N = 2$, $M = 1$). We shall use the
same notation as in section III. The variables to be determined are
$k_+$, $k_-$ and $ \Lambda$, where $k_+$ is the complex conjugate
of $k_-$ and $\Lambda$ is real. We may write
\beq
k_+ = q + \i \x \qd, \qd k_- = q - \i \x \qd,
\eeq
where $0 \le q < 2 \pi $ and $\x > 0$. Then the total momentum is
\beq
P = 2q = \frac {2 m \pi} L \mod 2 \pi \qd.
\eeq
This equation restricts the admissible values of $q$. It is easy to
obtain $\La$ as a function of $q$ and $\xi$. We find $\La = \sin(q)
\cosh(\x )$ (cf.\ equation (\ref{laq})). Thus we are left with a single
equation that determines $\xi$. For our numerical calculations we
wrote it in the form
\beq
\exp( \i m \pi + \x L ) = \frac{-\cos(m \p/L) \sinh(\xi) + U/4}
{-\cos(m \p/L) \sinh(\xi) - U/4}
\qd. \label{realeq}
\eeq
This equation is equivalent to (\ref{xiq}). Therefore, for positive $U$,
the allowed values of $m$ are restricted to
\beq
m = L/2 + 1, \ldots, 3L/2 - 1
\eeq
(cf.\ equation (\ref{rangem})). Equation (\ref{realeq}) was already
studied in appendix B of \cite{EKS92b}. There it was shown that there
is a redistribution phenomenon as $U$ becomes small. $k$-$\La$ strings
corresponding to odd values of $m$ collapse at critical values of
$U$ given by $U_m = (8/L)|\cos (m\pi/L)|$.
\subsubsection{Numerical solutions for $N=2$}
As a typical example, let us present some numerical $k$-$\La$ two
string solutions for $N=2$ (two electrons). For a $k$-$\La$ two string
we show the dependence of energy eigenvalues, imaginary parts of
charge momenta and the deviation from the ideal string positions on $U$.
We put $m=16$ and $L=16$ (16 sites). Then we have $q=\pi$, i.e.
$$
k_+ = \pi + \i \xi \qd, \qd k_- = \pi - \i \xi \qd,
\qd \La = 0 \qd.
$$
In the list below the deviation of the string from its ideal position,
$k_{\pm}^{(ideal)} = \pi \pm \i \, \arsh (U/4)$ is measured by
$\Im \delta = \sinh (\xi ) - U/4$.
\vskip 0.6cm
(1) $U=10$
\begin{eqnarray}
{\rm Householder} \qquad
E & = & 10.7703296143355 \nonumber \\
{\rm Bethe \, Ansatz} \qquad
E & = & 10.7703296143355 \nonumber \\
\x & = & 1.64723114637774 \nonumber \\
\Im \delta & = & 1.78994596922166 \times 10^{-11} \nonumber
\end{eqnarray}
(2) $U=1$
\begin{eqnarray}
{\rm Householder} \qquad
E & = & 4.13148449882288 \nonumber \\
{\rm Bethe \, Ansatz} \qquad
E & = & 4.13148449882289 \nonumber \\
\x &=& 0.255705305537532 \nonumber \\
\Im \delta & =& 8.50098694369150 \times 10^{-3} \nonumber
\end{eqnarray}
(3) $U=0.1$
\begin{eqnarray}
{\rm Householder} \qquad
E & = & 4.00668762410970 \nonumber \\
{\rm Bethe \, Ansatz} \qquad
E & = & 4.00668762410971 \nonumber \\
\x & = & 5.78176505356310 \times 10^{-2} \nonumber \\
\Im \delta &= & 3.28498688384022 \times 10^{-2} \nonumber
\end{eqnarray}
(4) $U=0.01$
\begin{eqnarray}
{\rm Householder} \qquad
E& = & 4.00062917225929 \nonumber \\
{\rm Bethe \, Ansatz} \qquad
E & = & 4.00062917225931 \nonumber \\
\x & = & 1.77363435624506 \times 10^{-2} \nonumber \\
\Im \delta &=& 1.52372734873120 \times 10^{-2} \nonumber
\end{eqnarray}
The string with $m=16$ does exist for any $U>0$ and is actually the
highest level in the spectrum for $N=2$. We note that the non-ideal
string approaches the ideal $k$-$\La$ string, when $U$ becomes large,
even for such a small system. Yet, in accordance with our expectations,
the ideal string does not provide a good approximation for small $U$ in
finite systems.
In table 1 we show a complete list of eigenstates for the case of one
down-spin ($M=1$, $S_z=0$) and $U=1.5$ for a 6-site system ($L=6$).
Note that the value of $U = 1.5$ is greater than $\max U_m=1.1547$,
or, in the language of section III, the Takahashi condition is
satisfied. Let us explain the table.
\begin{enumerate}
\item
The 36 eigenstates are listed in increasing order with respect to
their energy.
\item
$S$ and $P$ denote the spin and momentum of the eigenstate,
respectively.
\item
The energy eigenvalues obtained by direct diagonalization of the
Hamiltonian (the Householder method) and by the Bethe ansatz method
coincide within an error of $\CO(10^{-15})$.
\item
The last digit for each numerical value has a rounding error.
\item
There are 5 $k$-$\La$ string solutions among the 36 eigenstates, which
is consistent with the number $L-1 = 5$ obtained by Takahashi's counting
(\ref{r1})-(\ref{r3}).
\end{enumerate}
Let us give more explanations on the table. In fact, it confirms the
completeness of the Bethe ansatz as discussed in section II.D.
We first note that there are 36 eigenstates for the case of two
electrons and one down-spin ($N=2$ and $M=1$) on a 6-site lattice
($L=6$). We recall that the two electrons with one up-spin and one
down-spin can occupy the same site. The ${6 \choose 1} \times
{6 \choose 1} = 36$ eigenstates in table 1 can be classified
into the following types.
\begin{enumerate}
\item 15 eigenstates with two real charge momenta $k_1$, $k_2$
and one real spin-rapidity $\Lambda$.
\item 5 eigenstates with one $k$-$\Lambda$ two string.
\item 15 eigenstates with $S=1$ and $S^z=0$ belonging to spin triplets.
\item 1 $\h$-pairing state.
\end{enumerate}
Let us consider case (i). In table 1 there are 15 states with real
charge momenta which have total spin zero ($S=0$). This agrees
with our analytical arguments in section III on the number of real,
regular Bethe ansatz solutions (cf.\ Lemma III.2 and below), which
should be ${6 \choose 2} = 15$. Let us also recall that the $I_j$ are
integer (half-integer) when $\sum_m M_m + M_m^{'}$ is even (odd).
Thus, for the case $M=1$, the $I_j$ should be half-odd integer,
which is in agreement with the results shown in table 1.
We now consider case (ii). From the analytic discussion in section
III.C, we should have 5 eigenstates with $k$-$\La$ strings. This
is in accordance with our numerical data in table 1. Recall that
we showed in section III.C (below Lemma III.3) that also Takahashi's
counting, using equations (\ref{r1})-(\ref{r3}), leads to the same
number of $k$-$\La$ string solutions.
\begin{center}
\tabcolsep4mm
\begin{tabular}{|c|c|r|r|r|}
\hline
No. & Energy & $S$ & $P/(\p/3)$ & type of solution
\nonumber \\ \hline
1 & $-3.82047006625301$
& 0 & 0 & real $I_j = -0.5, 0.5$
\nonumber \\ \hline
2 & $-3.00000000000000$ & 1 & 1 & triplet $I_j = 0, 1$
\nonumber \\ \hline
3 & $-3.00000000000000$ & 1 & 5 & triplet $I_j = 0, -1$
\nonumber \\ \hline
4 & $-2.60959865138515$
& 0 & 1 & real $I_j = -0.5, 1.5$
\nonumber \\ \hline
5 & $-2.60959865138515$
& 0 & 5 & real $I_j = -1.5, 0.5$
\nonumber \\ \hline
6 & $-2.00000000000000$ & 1 & 0 & triplet $I_j = 1, -1$
\nonumber \\ \hline
7 & $-1.86256622153075$
& 0 & 2 & real $I_j = 0.5, 1.5$
\nonumber \\ \hline
8 & $-1.86256622153075$
& 0 & 4 & real $I_j = -0.5, -1.5$
\nonumber \\ \hline
9 & $-1.53909577258373$
& 0 & 0 & real $I_j = -1.5, 1.5$
\nonumber \\ \hline
10 & $-1.00000000000000$ & 1 & 2 & triplet $I_j = 0, 2$
\nonumber \\ \hline
11 & $-1.00000000000000$ & 1 & 4 & triplet $I_j = 0, -2$
\nonumber \\ \hline
12 & $-0.594210897273940$
& 0 & 2 & real $I_j = -0.5, 2.5$
\nonumber \\ \hline
13 & $-0.594210897273940$
& 0 & 4 & real $I_j = -2.5, 0.5$
\nonumber \\ \hline
14 & 0.00000000000000 & 0 & 3 & real $I_j = 0.5, 2.5$
\nonumber \\ \hline
15 & 0.00000000000000 & 0 & 3 & real $I_j = -2.5, -0.5$
\nonumber \\ \hline
16 & 0.00000000000000 & 1 & 3 & triplet $I_j = 1, 2$
\nonumber \\ \hline
17 & 0.00000000000000 & 1 & 3 & triplet $I_j = -1, -2$
\nonumber \\ \hline
18 & 0.00000000000000 & 1 & 3 & triplet $I_j = 0, 3$
\nonumber \\ \hline
19 & 0.00000000000000 & 1 & 1 & triplet $I_j = 2, -1$
\nonumber \\ \hline
20 & 0.00000000000000 & 1 & 5 & triplet $I_j = -2, 1$
\nonumber \\ \hline
21 & 0.474357244982949
& 0 & 1 & real $I_j = -1.5, 2.5$
\nonumber \\ \hline
22 & 0.474357244982949
& 0 & 5 & real $I_j = -2.5, 1.5$
\nonumber \\ \hline
23 & 1.00000000000000 & 1 & 2 & triplet $I_j = 3, -1$
\nonumber \\ \hline
24 & 1.00000000000000 & 1 & 4 & triplet $I_j =3, 1$
\nonumber \\ \hline
25 & 1.43079477929458
& 0 & 4 & real $I_j = 1.5, 2.5$
\nonumber \\ \hline
26 & 1.43079477929458
& 0 & 2 & real $I_j = -2.5, -1.5$
\nonumber \\ \hline
27 & 1.50000000000000 & 0 & 3 & $\h$-pair
\nonumber \\ \hline
28 & 2.00000000000000 & 1 & 0 & triplet $I_j = 2, -2$
\nonumber \\ \hline
29 & 2.49213401737360
& 0 & 0 & real $I_j =-2.5, 2.5$
\nonumber \\ \hline
30 & 2.52598233951011
& 0 & 2 & complex $m=8$, $\xi = 0.710224864788777$
\nonumber \\ \hline
31 & 2.52598233951011
& 0 & 4 & complex $m=4$, $\xi = 0.710224864788777$
\nonumber \\ \hline
32 & 3.00000000000000 & 1 & 5 & triplet $I_j = 2, 3$
\nonumber \\ \hline
33 & 3.00000000000000 & 1 & 1 & triplet $I_j = -2, 3$
\nonumber \\ \hline
34 & 3.63524140640220
& 0 & 1 & complex $m=7$, $\xi = 0.313056827256169$
\nonumber \\ \hline
35 & 3.63524140640220
& 0 & 5 & complex $m=5$, $\xi = 0.313056827256169$
\nonumber \\ \hline
36 & 4.36743182146314
& 0 & 0 & complex $m=6$, $\xi = 0.425405934759021$
\nonumber \\ \hline
\end{tabular}
\vskip 0.6cm
{Table 1: Classification of all energy levels for $L = 6$, $N = 2$,
$M = 1$ and $U = 1.5$.}
\end{center}
We consider case (iii). There are 15 states with $S=1$ and $S_z=0$.
We describe them as {\it triplets} in table 1. They are obtained by
multiplying the spin-lowering operator $\zeta^{\dagger}$ to the regular
Bethe states with $S=1$ and $S_z=1$. (For the notation for the $SO(4)$
symmetry see section II.B). The regular Bethe states with $S=1$ and
$S_z=1$ correspond to $N=2$ and $M=0$. We recall again that the $I_j$'s
are integer (half-integer) when $\sum_m M_m + M_m^{'}$ is even (odd).
Thus, the $I_j$'s which belong to the $S^z = 0$ states in spin triplets
should be integer-valued. They should take one of the values $-2$,
$-1,$ $0$, $1$, $2$, or $3$, which means that there are ${6 \choose 2}
= 15$ states according to Takahashi's counting (\ref{r1})-(\ref{r3}).
There is one $\h$-pairing state. The energy of this state is equal to
$U$, since the two electrons occupy the same site.
Now, let us sum up all the numbers of the different types of
eigenstates:
\beq
15 + 5 + 15 + 1 = 36 \quad .
\eeq
Thus, we have shown that all the energy eigenstates obtained by Method
1 are confirmed by Method 2. In particular, we have confirmed
numerically the completeness of the Bethe ansatz.
Let us consider the total momentum. Using equations (\ref{t1})-%
(\ref{t3}) we can express the total momentum $P$ of the eigenstates
with real charge momenta in terms of $I_1$ and $I_2$,
\beq
P = {\frac {2 \pi} L} (I_1 + I_2) \; \mod 2 \pi.
\eeq
This formula is consistent with table 1.
Let us now discuss the $U$-dependence of the spectrum. In figure 4
we show the spectral flow from strong-coupling to weak-coupling, where
the Takahashi condition does not hold.
\begin{figure}
\begin{center}
\includegraphics[width = 8cm,angle = 270]{eneflow.eps}
\caption{The spectral flows for $N = 2$ and $M = 1$ for a 6-site
lattice. Solid lines denote the $k$-$\Lambda$ strings. Dashed lines
denote the energy of real roots ($S=0$). Dotted lines denote the
triplet states ($S=1$). The dash-dotted line denotes the energy of the
$\h$-pair. At $U_5=U_7=1.1547$, as indicated by the arrow,
two $k$-$\Lambda$ two strings with $m=5,7$ collapse into real
solutions.}
\end{center}
\end{figure}
In figure 4 we show the redistribution phenomenon discussed in
sections II.D, III.A and above. There are five $k$-$\La$ two-string
solutions in table 1. The entries 30, 31 and 36 correspond to even $m$.
According to our discussion above, these $k$-$\La$ two strings are
stable as $U$ becomes small. Entry number 36 is the highest energy
level in the figure. Entries number 30 and 31 are degenerate. They
correspond to the third highest level at $U = 5$. The entries number
34 and 35 correspond to odd $m = 5, 7$. The corresponding levels
are degenerate. At $U_5 = U_7 =1.1547$ these $k$-$\La$ two-string
solutions collapse into pairs of real charge momenta. This is indicated
by the arrow in figure 4. On the other hand, all five $k$-$\Lambda$
string solutions do exist as long as the Takahashi condition is
satisfied.
\subsection{Numerical solution for three electrons}
\subsubsection{Equations for the $k$-$\Lambda$ string}
We shall now consider the set of solutions to the Lieb-Wu equations
for $N=3$. The $k$-$\La$ strings to be searched for have two complex
$k_j$, a real $k_j$ and a real $\La$. We express these momenta and
rapidities by
\beq
k_1 = q - \i \xi \qd, \qd k_2 = q + \i \xi \qd,
\qd k_3 \qd, \qd \La \qd.
\eeq
The total momentum is $2 \pi m/L$, so that
\beq
k_1 + k_2 + k_3 = 2 \pi m/L \qd, \qd 2 \pi (\ell - 1)/L < k_3
< 2\pi \ell /L \qd.
\eeq
We define $\Re \delta$ and $\Im \delta$ by
\begin{eqnarray}
\sin k_1 & = & \Lambda + \i U/4 + \Re \delta + \i \, \Im \delta
\qd, \\
\sin k_2 & = & \Lambda - \i U/4 + \Re \delta - \i \, \Im \delta
\qd.
\end{eqnarray}
We recall that $\Re \delta$ and $\Im \delta$ describe the deviations
from the ideal string solutions. Now we derive equations for four
variables, $k_3$, $q$, $\x$ and $\La$, starting from the Lieb-Wu
equations (\ref{bak3}), (\ref{bas3}). Taking logarithms, we obtain a
set of equations of the form $f_i=0$ ($i=1,\ldots 4$). Here the $f_i$'s
are defined by
\begin{eqnarray}
f_1 & = & L k_3
- 2 \arctan \left( \frac{\Lambda - \sin k_3}{U/4} \right)
- 2 \pi(\ell - 1/2) \qd, \\
f_2 & = & - 2 L \arcsin
\bigg(1/2 \Big\{ \sqrt{(\Lambda + \Re \delta + 1)^2 +
(U/4 + \Im \delta )^2} \nonumber\\ \\
&& \qqqd - \sqrt{(\Lambda + \Re \delta -1)^2
+ (U/4 + \Im \delta)^2} \Big\} \bigg) +
L k_3 - 2 \pi(\ell - 1/2 - J') \qd, \\
f_3 & = & \exp(- 2 L \xi)- \frac{(\Re \delta)^2 + (\Im \delta)^2}
{(\Re \delta)^2 + (U/2 + \Im
\delta)^2} \qd, \\
f_4 & = & - 2 \arctan \left( \frac{\Re \delta}{\Im \delta}
\right)
+ 2 \arctan \left( \frac{\Re \delta}{U/2 +
\Im \delta} \right)
- 2 \arctan \left( \frac{\Lambda - \sin k_3}{U/4}
\right) - \sigma \pi \qd,
\end{eqnarray}
where
\beq
\Re \delta = \sin q \cosh \xi - \Lambda \qd, \qd
\Im \delta = - U/4 - \cos q \sinh \xi \qd,
\eeq
and the parameter $\sigma$ is given by
\beq
\sigma = 1 \qd {\rm for} \qd k_3 <
\frac{2 \pi}{L} (\ell - 1/2) \qd, \qd
\sigma = - 1 \qd {\rm for} \qd k_3 >
\frac {2 \pi}{L} (\ell - 1/2) \qd.
\eeq
Note that the equation $f_1 = 0$ is equivalent to equation (\ref{laofk})
for $j = 3$.
To solve this set of coupled equations by Method 2, we need a
proper initial guess. We employ the ideal strings given by
the discrete Takahashi equations as initial approximation.
The fact that the ideal strings provide a good estimate for
the true solution is crucial to reach the correct answer.
This is because the $f_i$'s are rather singular functions having many
diverging points. Very often, the true solution is very close to a
divergent point. We cannot approach the solution from a point beyond
a branch cut using an iterative way like the Brent method.
\subsubsection{Numerical examples of $k$-$\Lambda$ string solutions}
We present some numerical solutions for one $k$-$\La$ two string
and one real $k$. The parameters in the examples below are $N=3$
(three electrons), $L=10$ (10 sites) and $U=5$.
\vskip 0.6cm
(1) $m=10$, $\ell=1$ ($I_3=0, J^{'}=1/2$)
\begin{eqnarray}
{\rm Householder} \qquad
E & = & 4.46666961980768 \nonumber \\
{\rm Bethe \, Ansatz} \qquad
E & = & 4.46666961980768 \nonumber \\
q& = & 2.98893848049280 \nonumber \\
\xi & =& 1.05674954466496 \nonumber \\
k_3 & =& 0.305308346193987 \nonumber \\
\Lambda & =& 0.245232889885225 \nonumber \\
\Re \delta & =& -6.42851441900183 \times 10^{-5} \nonumber \\
\Im \delta & = & 2.84511548476196 \times 10^{-6} \nonumber
\end{eqnarray}
(2) $m=7$, $\ell=1$ ($I_3=0, J^{'}=7/2$)
\begin{eqnarray}
{\rm Householder} \qquad E & = & 3.48250616148668 \nonumber \\
{\rm Bethe \, ansatz} \qquad
E & = & 3.48250616148668 \nonumber \\
q& = & 1.92454676428806 \nonumber \\
\xi & = & 1.99506896270533 \nonumber \\
k_3 & = & 0.549136186449599 \nonumber \\
\Lambda & = & 3.51250692222577 \nonumber \\
\Re \delta & = & 2.08765360554253 \times 10^{-9} \nonumber \\
\Im \delta & = & 4.99459629210719 \times 10^{-9} \nonumber
\end{eqnarray}
Let us discuss possible numerical errors for the above solutions.
Their numerical errors may be evaluated by the residual, $f_i$,
given for these solutions as follows.
{\vskip 0.6cm }
(1) $m=10$, $\ell=1$ ($I_3=0, J=1/2$)
\begin{eqnarray}
f_1 & = & -1.24900090270330 \times 10^{-15} \nonumber \\
f_2 & = & -3.99680288865056 \times 10^{-15} \nonumber \\
f_3 & = & 5.25259688971173 \times 10^{-22} \nonumber \\
f_4 & = & 4.53592718940854 \times 10^{-12} \nonumber
\end{eqnarray}
with $\sigma=+1$.
(2) $m=7$, $\ell=1$ ($I_3=0, J=7/2$)
\begin{eqnarray}
f_1 & =& -1.33226762955019 \times 10^{-15} \nonumber \\
f_2 & =& 1.42108547152020 \times 10^{-15} \nonumber \\
f_3 & =& -4.66698631842825 \times 10^{-25} \nonumber \\
f_4 & =& -7.67983898697366 \times 10^{-10} \nonumber
\end{eqnarray}
with $\sigma=-1$.
In comparison to other error values $f_i$ the number $f_4$ has a
rather large value. However, in $f_4$, we have an expression like
${\rm e}/{\rm e}'$ (${\rm e} \simeq 0$, ${\rm e}' \simeq 0$). So this value contains a
larger cancellation error for smaller $\Re \delta$ and $\Im \delta$.
Note again that relative errors for the energy are always
$\CO(10^{-15})$.
\subsubsection{A complete list of solutions for $N=3$ and $M=1$}
We can present complete lists of eigenstates for all finite systems
tractable by our numerical technique. As a further example, we consider
all eigenstates for $N=3$ and $M=1$ in a 6-site system ($L=6$). The
list is shown in appendix~C. It confirms again the completeness of
the Bethe ansatz.
Let us briefly discuss the numbers of eigenstates of different types.
First, we note that there are in total $6 {6 \choose 2} = 90$
eigenstates. Inspection of the tables in appendix~C shows that they
can be classified into the following four types.
\begin{enumerate}
\item
40 eigenstates with three real charge momenta $k_1, k_2, k_3$ and
one real spin rapidity $\Lambda$.
\item
24 eigenstates with one $k$-$\Lambda$ string with $k_1 = k_- = q - \i
\xi$, $k_2 = k_+ = q + \i \xi$, and $\Lambda$ and $k_3$ real.
\item
20 eigenstates with $S = 3/2$ and $S_z = 1/2$ belonging to spin
quartets.
\item
6 eigenstates with one $\h$-pair and one real charge momentum.
\end{enumerate}
Let us now confirm that these numbers agree with Takahashi's counting,
(\ref{r1})-(\ref{r3}): Case (i) was considered in section III.A below
Lemma III.2. There we showed that Takahashi's counting predicts a
number of $2 {L \choose 3}$ real solutions for three electrons and
one down spin. Inserting $L = 6$ we have $2 {6 \choose 3} = 40$
eigenstates, which is in accordance with our numerical calculation.
Case (ii) was considered below Lemma III.5. The number of eigenstates
obtained there by Takahashi's counting was $L(L - 2)$, which for
$L = 6$ gives as desired $6 \cdot 4 = 24$. Let us consider case (iii).
These states are the second highest states ($S_z = 1/2$) in spin
quartets. They are obtained from regular Bethe states with $N = 3$,
$M = 0$ by multiplication with the spin lowering operator $\z^\dagger$.
Since $N = 3$ and $M = 0$, we have ${6 \choose 3} = 20$ states of this
type. The $\h$-pair in case (iv) is obtained by acting with $\h^+$ on
regular Bethe states with $N = 1$ and $M = 0$. Hence, Takahashi's
counting gives 6 eigenstates of this type on a 6-site lattice.
\subsection{Summary}
In this section we have presented a thorough numerical study of the
Hubbard model. We calculated, in particular, all eigenstates and
eigenvectors for a six-site lattice with two and three electrons and
one down spin by direct numerical diagonalization of the Hamiltonian.
These data were compared with data obtained by numerical solution of
the Lieb-Wu equations. Both sets of data are in perfect numerical
agreement and confirm once again the results of our analytical
investigation in the previous sections. The structure of our numerical
data is fully consistent with Takahashi's string hypothesis. The
number and classification of the eigenstates is consistent with
Takahashi's counting (\ref{r1})-(\ref{r3}). Thus, our numerical data
confirm not only the existence of $k$-$\Lambda$ strings but also the
completeness of the Bethe ansatz. Without $k$-$\La$ strings the Bethe
ansatz would be incomplete.
@
\section{Thermodynamics in the Yang-Yang approach and excitation
spectrum in the infinite volume}
Let us now turn to the determination of thermodynamic quantities
and the zero-temperature excitation spectrum in the
infinite volume. A convenient way to construct the spectrum was
pioneered by C.N. Yang and C.P. Yang for the case of the
delta-function Bose gas \cite{YaYa69}. The starting point are
the Bethe Ansatz equations in the finite volume. They are used to
derive a set of coupled, nonlinear integral equations called
thermodynamic Bethe Ansatz (TBA) equations, which describe
the thermodynamics of the model at finite temperatures.
The quantities entering these equations have a natural interpretation
in terms of dressed energies of elementary excitations.
Yang and Yang's formalism is a natural generalization of the
thermodynamics of the free Fermi gas to interacting systems.
In what follows we review Takahashi's derivation of the TBA equations
for the case of the repulsive Hubbard model \cite{Takahashi72}. The
analogous calculations for the attractive case can be found in
\cite{LeSc89}.
Our starting point are the discrete Takahashi equations \r{t1}-\r{t3}
and expressions for energy \r{en} and momentum \r{mom} for very large
but finite $L$. A very important property of \r{t1}-\r{t3} is that as we
approach the thermodynamic limit $L\to \infty$, $N/L$ and $M/L$ fixed
(finite densities of electrons and spin down electrons), the roots of
\r{t1}-\r{t3} become dense
\begin{equation}
k_{j+1}-k_j={\cal O}(L^{-1}),\quad
\Lambda^n_{\alpha+1}-\Lambda^n_\alpha={\cal O}(L^{-1}),\quad
{\Lambda^\prime}^n_{\alpha+1}-{\Lambda^\prime}^n_\alpha=
{\cal O}(L^{-1}).
\label{dense}
\end{equation}
We now define so-called {\sl counting functions}
$y,z_n,z^\prime_n$ as follows
\begin{eqnarray}
k L & = & L y(k) - \sum_{n=1}^\infty \sum_{\alpha = 1}^{M_n}
\theta \left(
\frac{\sin k - \Lambda_\alpha^n}{nU/4} \right)
- \sum_{n=1}^\infty \sum_{\alpha = 1}^{M_n'}
\theta \left(
\frac{\sin k - {\Lambda'}_\alpha^n}{nU/4} \right),
\\
\sum_{j=1}^{N - 2M'} \theta \left(
\frac{\Lambda - \sin k_j}{nU/4} \right) & = &
L z_n(\Lambda) +
\sum_{m=1}^\infty \sum_{\beta = 1}^{M_m}
\Theta_{nm} \left(
\frac{\Lambda - \Lambda_\beta^m}{U/4} \right),
\\
L [\arcsin({\Lambda'} + n \tst{\frac{\i U}{4}})
+ \arcsin({\Lambda'} - n \tst{\frac{\i U}{4}})] & = &
L z^\prime_n(\Lambda') +
\sum_{j=1}^{N - 2M'} \theta \left(
\frac{{\Lambda'} - \sin k_j}{nU/4} \right) +
\sum_{m=1}^\infty \sum_{\beta = 1}^{M_m'}
\Theta_{nm} \left(
\frac{{\Lambda'} - {\Lambda'}_\beta^m}{U/4}
\right).
\end{eqnarray}
By definition the counting functions satisfy the following equations
when evaluated for a given solution of the discrete Takahashi equations
\begin{equation}
y(k_j)=2\pi I_j/L\ ,\quad
z^\prime_n({\Lambda^\prime}^n_\alpha)=2\pi {J^\prime}^n_\alpha/L\ ,\quad
z_n({\Lambda}^n_\alpha)=2\pi J^n_\alpha/L\ .
\end{equation}
In the next step we define the so-called {\sl root densities},
which are related to the counting functions as follows. By definition
the counting functions ``enumerate'' the Bethe Ansatz roots e.g.
\begin{equation}
L[y(k_j)-y(k_n)]=2\pi(I_j-I_n).
\end{equation}
For a given solution of \r{t1}-\r{t3} certain of the (half-odd)
integers between $I_j$ and $I_n$ will be ``occupied'' i.e. there will
be a corresponding root $k$, whereas others will be omitted. We describe
the corresponding $k$-values in terms of a root density $\rho(k)$ for
``particles'' and a density $\rho^h(k)$ for ``holes''. In a very large
system we then have by definition (here the property \r{dense} is very
important)
\bea
L\rho(k)\ dk&=& {\rm number\ of\ k's\ in\ dk}\ ,\nonumber\\
L\rho^h(k)\ dk&=& {\rm number\ of\ holes\ in\ dk}\ .
\eea
It is then clear that in the thermodynamic limit we have
\begin{equation}
2\pi[\rho(k)+\rho^h(k)]=\frac{dy(k)}{dk}\ .
\end{equation}
The analogous equations for the other roots of \r{t1}-\r{t3} are
\bea
2\pi[\sigma_n(\Lambda) + \sigma_n^h(\Lambda)] =
\frac{dz_n(\Lambda)}{d\Lambda}\ , \qquad
2\pi[\sigma^\prime_n(\Lambda)+{\sigma_n^\prime}^h(\Lambda)]=
\frac{dz^\prime_n(\Lambda)}{d\Lambda}\ .
\eea
In the thermodynamic limit the discrete Takahashi equations can
now be turned into coupled integral equations involving both counting
functions and root densities
\begin{eqnarray} \label{z1}
k & = & y(k) - \sum_{n=1}^\infty \int_{-\infty}^\infty
d\Lambda\ \theta \left(
\frac{\sin k - \Lambda}{nU/4} \right)
\left[\sigma^\prime_n(\Lambda)+\sigma_n(\Lambda)\right],
\\ \label{z2}
\int_{-\pi}^\pi dk\ \theta \left(
\frac{\Lambda - \sin k}{nU/4} \right)\ \rho(k) & = &
z_n(\Lambda) +
\sum_{m=1}^\infty \int_{-\infty}^\infty
d\Lambda^\prime\ \Theta_{nm} \left(
\frac{\Lambda - \Lambda^\prime}{U/4} \right)\
\sigma_m(\Lambda^\prime),
\\ \label{z3}
\arcsin({\Lambda} + n \tst{\frac{\i U}{4}})
+ \arcsin({\Lambda} - n \tst{\frac{\i U}{4}}) & = &
{z^\prime_n}(\Lambda) +
\int_{-\pi}^\pi dk\ \theta \left(
\frac{{\Lambda} - \sin k}{nU/4} \right)\ \rho(k)\nonumber\\
&& +
\sum_{m=1}^\infty \int_{-\infty}^\infty d\Lambda^\prime\
\Theta_{nm} \left(
\frac{{\Lambda} - {\Lambda^\prime}}{U/4}
\right)\ \sigma^\prime_m(\Lambda^\prime).
\end{eqnarray}
As we are interested in the Hubbard model at finite temperatures we
need to express the entropy in terms of the root densities. This is
achieved by observing that e.g. the number of vacancies for $k$'s in
the interval $[k,k+dk]$ is simply $L(\rho(k)+\rho^h(k))dk$. Of these
``vacancies'' $L\rho(k)dk$ are occupied. The corresponding
contribution $dS$ to the entropy is thus formally
\begin{equation}
e^{dS}=\frac{(L[\rho(k)+\rho^h(k)]dk)!}{(L\rho(k)dk)!(L\rho^h(k)dk)!}\ ,
\end{equation}
where $!$ denotes the factorial. The differential $dS$ is
obtained via Stirling's formula. After integration we obtain the
following expression for the total entropy density of the Hubbard
model
\bea
S/L&=&\int_{-\pi}^\pi dk\left\lbrace
[\rho(k)+\rho^h(k)] \ln[\rho(k)+\rho^h(k)]
- \rho(k) \ln\rho(k) - \rho^h(k) \ln\rho^h(k)
\right\rbrace\nonumber\\
&&+\sum_{n=1}^\infty\int_{-\infty}^\infty d\Lambda\left\lbrace
[\sigma_n(\Lambda)+\sigma_n^h(\Lambda)]
\ln[\sigma_n(\Lambda)+\sigma_n^h(\Lambda)]-
\sigma_n(\Lambda)\ln\sigma_n(\Lambda) -
\sigma_n^h(\Lambda)\ln\sigma_n^h(\Lambda)
\right\rbrace\nonumber\\
&&+\sum_{n=1}^\infty\int_{-\infty}^\infty d\Lambda\left\lbrace
[\sigma^\prime_n(\Lambda)+{\sigma_n^\prime}^h(\Lambda)]
\ln[\sigma_n^\prime (\Lambda)+{\sigma^\prime_n}^h(\Lambda)]-
\sigma^\prime_n(\Lambda)\ln{\sigma^\prime_n}(\Lambda) -
{\sigma^\prime_n}^h(\Lambda)\ln{\sigma^\prime_n}^h(\Lambda)
\right\rbrace .
\eea
The Gibbs free energy per site is thus
\bea
f&=&\frac{E-\mu N-2BS^z-TS}{L}\nonumber\\
&=&\int_{-\pi}^\pi dk\left[-2\cos k -\mu-U/2-B\right]\rho(k)
+\sum_{n=1}^\infty\int_{-\infty}^\infty d\Lambda\ 2nB\ \sigma_n(\Lambda)\nonumber\\
&&+\sum_{n=1}^\infty\int_{-\infty}^\infty d\Lambda\left[ 4{\rm
Re}\sqrt{1-(\Lambda-inU/4)^2}-2n\mu-nU\right]\sigma^\prime_n(\Lambda)-T\ S/L\ .
\label{gibbs1}
\eea
Here $\mu$ is a chemical potential, $B$ is a magnetic field and $T$ is
the temperature.
The alert reader will have realized that we are still missing
a set of equations that allows us to completely determine the root
densities and counting functions (the thermodynamic limit
\r{z1}-\r{z3} of the discrete Takahashi equations are clearly
insufficient). This is the topic of the following subsection.
\subsection{Takahashi's Thermodynamic Equations}
Let us start by differentiating \r{z1}-\r{z3}, which yields
\bea
\rho(k)+\rho^h(k)&=&\frac{1}{2\pi}+\cos
k\sum_{n=1}^\infty\int_{-\infty}^\infty d\Lambda\frac{nU/4}{\pi}
\frac{\sigma^\prime_n(\Lambda)+\sigma_n(\Lambda)}{(nU/4)^2+(\sin
k-\Lambda)^2 }\ ,\nonumber\\
\sigma_n^h(\Lambda)&=&-\sum_{m=1}^\infty A_{nm}*\sigma_m\bigg|_\Lambda
+\int_{-\pi}^\pi dk \frac{nU/4}{\pi}\frac{\rho(k)}{(nU/4)^2+(\sin
k-\Lambda)^2 }\ ,\nonumber\\
{\sigma^\prime_n}^h(\Lambda)&=&\frac{1}{\pi}{\rm
Re}\frac{1}{\sqrt{1-(\Lambda -inU/4)^2}}
-\sum_{m=1}^\infty A_{nm}*\sigma^\prime_m\bigg|_\Lambda
-\int_{-\pi}^\pi dk \frac{nU/4}{\pi}\frac{\rho(k)}{(nU/4)^2+(\sin
k-\Lambda)^2 }\ .
\label{densities}
\eea
Here $A_{nm}$ is an integral operator
acting on a function $f$ as
\bea
A_{nm}*f\bigg|_x&=&\delta_{nm}f(x)+\frac{d}{dx}\int_{-\infty}^\infty
\frac{dy}{2\pi} \Theta_{nm}\left(\frac{x-y}{U/4}\right)f(y).
\eea
Equations \r{densities} can be used to express the densities of holes
in terms of densities of particles.
A second set of equations is obtained by considering the Gibbs free
energy density \r{gibbs1} as a functional of the root densities. In
thermal equilibrium $f$ is stationary with respect to variations
of the root densities
\bea
0&=&\delta f \nonumber\\
&=&\frac{\delta f}{\delta\rho(k)}\delta\rho(k)
+\frac{\delta f}{\delta\rho^h(k)}\delta\rho^h(k)
+\sum_{n=1}^\infty\left[\frac{\delta
f}{\delta\sigma_n(\Lambda)}\delta\sigma_n(\Lambda)
+\frac{\delta f}{\delta\sigma^h_n(\Lambda)}\delta\sigma^h_n(\Lambda)
+\frac{\delta
f}{\delta{\sigma^\prime_n}(\Lambda')}\delta{\sigma^\prime_n}(\Lambda')
+\frac{\delta f}{\delta{\sigma^\prime_n}^h(\Lambda')}
\delta{\sigma^\prime_n}^h(\Lambda')\right],\nonumber\\
\eea
where we need to take into account \r{densities} as constraint
equations. In this way one obtains a set of equations for the ratios
$\zeta=\rho^h/\rho$, $\eta_n=\sigma_n^h/\sigma_n$ and
$\eta_n^\prime={\sigma^\prime_n}^h/\sigma_n^\prime$
\bea
&&\ln \zeta(k)=\frac{-2\cos k -\mu-U/2-B}{T}+\sum_{n=1}^\infty
\int_{-\infty}^\infty\!\! d\Lambda \frac{nU}{4\pi}
\frac{1}{(nU/4)^2+(\sin k-\Lambda)^2}\left[
\ln\left(1+\frac{1}{\eta^\prime_n(\Lambda)}\right)
-\ln\left(1+\frac{1}{\eta_n(\Lambda)}\right)\right].\nonumber\\
\label{zeta}
\eea
\bea
&&\ln\left(1+\eta_n(\Lambda)\right) +\int_{-\pi}^\pi dk \frac{\cos k}{\pi}
\frac{nU/4}{(nU/4)^2+(\sin k-\Lambda)^2}\
\ln\left(1+\frac{1}{\zeta(k)}\right)=\frac{2nB}{T}+
\sum_{m=1}^\infty A_{nm}*\ln\left(1+\frac{1}{\eta_m}\right)\bigg|_\Lambda.
\label{lambda}
\eea
\bea
&&\ln\left(1+\eta^\prime_n(\Lambda)\right)
+\int_{-\pi}^\pi dk \frac{\cos k}{\pi}
\frac{nU/4}{(nU/4)^2+(\sin k-\Lambda)^2}\
\ln\left(1+\frac{1}{\zeta(k)}\right)=\nonumber\\
&&=\frac{4{\rm Re}\sqrt{1-(\Lambda -inU/4)^2}-2n\mu-nU}{T}
+\sum_{m=1}^\infty A_{nm}*\ln\left(1+\frac{1}{\eta^\prime_m}\right)
\bigg|_\Lambda.
\label{kl}
\eea
Note that \r{densities} together
with \r{zeta}-\r{kl} completely determine the densities of holes {\sl
and} particles in the state of thermal equilibrium.
The Gibbs free energy per site is given in terms of solutions of
\r{zeta}-\r{kl} as
\bea
f&=&-T\int_{-\pi}^\pi\frac{dk}{2\pi}\ \ln\left(1+\frac{1}{\zeta(k)}\right)
-T\sum_{n=1}^\infty\int_{-\infty}^\infty\frac{d\Lambda}{\pi}
\ \ln\left(1+\frac{1}{\eta^\prime_n(\Lambda)}\right) {\rm Re}\frac{1}{\sqrt{1-(\Lambda
-inU/4)^2}}\ .
\label{gibbs}
\eea
Following Takahashi we define
\begin{equation}
\kappa(k)=T\ln(\zeta(k))\ ,
\epsilon_n(\Lambda)=T\ln(\eta_n(\Lambda))\ ,
\epsilon^\prime_n(\Lambda)=T\ln(\eta^\prime_n(\Lambda))\ .
\end{equation}
As was first shown by Yang and Yang for the delta-function Bose
gas \cite{YaYa69}, the quantities defined in this way describe the
dressed energies of elementary excitations in the zero temperature
limit. Before turning to this we will give a brief summary on how to
calculate thermodynamic quantities in the framework of Takahashi's
approach.
\subsection{Thermodynamics}
The expression for the Gibbs free energy density \r{gibbs} can be simplified
\cite{Takahashi74}
\bea
f&=&E_0-\mu-U/2-T\left[\int_{-\pi}^\pi dk\ \rho_0(k)\
\ln\left(1+\zeta(k)\right)+\int_{-\infty}^\infty d\Lambda\ \sigma_0(\Lambda)
\ \ln\left(1+\eta_1(\Lambda)\right)
\right],
\label{gibbs2}
\eea
where
\bea
\sigma_0(\Lambda)&=&\int_{-\pi}^\pi dk \frac{1}{U}
\frac{1}{\cosh\frac{2\pi}{U}(\Lambda-\sin k)}\rho_0(k)\ ,\nonumber\\
\rho_0(k)&=&\frac{1}{2\pi}+\cos
k\int_{-\infty}^\infty \frac{d\Lambda}{\pi}
\frac{U/4}{(U/4)^2+(\sin k-\Lambda)^2 }\ \sigma_0(\Lambda)\ ,\nonumber\\
E_0&=&-2\int_{-\pi}^\pi dk\ \cos(k)\ \rho_0(k)
=-4\int_0^\infty d\omega\frac{J_0(\omega)J_1(\omega)}{1+\exp(\omega U/2)}\ .
\label{omega_simp}
\eea
We note that $\rho_0$, $\sigma_0$ and $E_0$ are the root density
for real $k$'s, the root density for real $\Lambda$'s and the ground
state energy for the half-filled repulsive Hubbard model, respectively.
Since the occurrence of quantities related to the half-filled Hubbard
model in \r{gibbs2} may be surprising, we would like to emphasize
that \r{gibbs2}, \r{omega_simp} holds for all negative values of the
chemical potential $\mu$ i.e.\ for all particle densities between zero
and one.
The representation \r{gibbs2} is convenient as it shows that the Gibbs
free energy is determined by the dressed energies for real $k$'s and
real $\Lambda$'s only. In order to derive \r{gibbs2} the following
identities are useful
\bea
&&\int_{-\pi}^\pi dk \frac{nU}{4\pi}\frac{\rho_0(k)}{(nU/4)^2+(\sin k-\Lambda)^2}
=\frac{1}{\pi}{\rm Re}\frac{1}{\sqrt{1-(\Lambda-inU/4)^2 }}\ ,\nonumber\\
&&\int_{-\pi}^\pi dk \frac{nU}{4\pi}\frac{\rho_0(k)}{(nU/4)^2+(\sin k-\Lambda)^2}
=A_{n1}*\sigma_0\bigg|_\Lambda\ .
\eea
At very low temperatures $T\ll B$ it is possible to determine the
Gibbs free energy by using an expansion of the TBA equations
\r{zeta}-\r{kl} for small $T$ \cite{Takahashi74}. The TBA equations
essentially reduce two only two coupled equations for $\zeta$ and
$\eta_1$ in this limit. For generic values of $B$ and arbitrary
temperatures one needs to resort to a numerical solution of
\r{zeta}-\r{kl}. In order to do so, one needs to truncate the
infinite towers of equations for $\Lambda$ and $k$-$\Lambda$ strings
at some finite value of their respective lengths.
In \cite{KUO89,UKO90} such a truncated set of equations was solved by
iteration. The integrals were discretized by using of the order of
$50$ ($100$) points for the $k$ ($\Lambda$) integrations.
The results of these computations are compared to the results of the
Quantum Transfer Matrix approach in the next section.
\subsection{Zero Temperature Limit}
For the remainder of this section we set the magnetic field to zero
$B=0$.
For $T\to 0$ the thermodynamic equations \r{zeta}-\r{kl} then reduce
to \cite{Takahashi72}
\bea
\kappa(k)&=&-2\cos k-\mu-U/2+\int_{-\infty}^\infty
\frac{d\Lambda}{\pi} \frac{U/4}{(U/4)^2+(\sin
k-\Lambda)^2}\epsilon_1(\Lambda)\ ,
\label{kappa}
\eea
\bea
(1-\delta_{n,1})\epsilon_n(\Lambda)&=&\int_{-Q}^Qdk\ \frac{\cos
k}{\pi} \frac{nU/4}{(nU/4)^2+(\sin k-\Lambda)^2}\ \kappa(k)
-A_{n1}*\epsilon_1\bigg|_\Lambda
\ ,\label{eps}
\eea
\bea
\epsilon_n^\prime(\Lambda)&=&
{4{\rm Re}\sqrt{1-(\Lambda -inU/4)^2}-2n\mu-nU}+\int_{-Q}^Qdk\
\frac{\cos k}{\pi} \frac{nU/4}{(nU/4)^2+(\sin k-\Lambda)^2}\ \kappa(k)
\ . \label{epsprime}
\eea
Note that $\kappa(\pm Q)=0$, $\kappa(k)<0$ for $|k|<Q$ and
$\epsilon_1(\Lambda)<0$. Using Fourier transform, equations \r{kappa}
and \r{eps} can be simplified further with the result
\bea
\kappa(k)&=&-2\cos k-\mu-U/2+\int_{-Q}^Qdk^\prime \cos k^\prime\
R(\sin k^\prime -\sin k) \ \kappa(k^\prime)\ ,\nonumber\\
\epsilon_1(\Lambda)&=&\int_{-Q}^Qdk \frac{\cos k}{U}
\frac{1}{\cosh\frac{2\pi}{U}(\Lambda-\sin k)}\kappa(k)\ ,\nonumber\\
\epsilon_n(\Lambda)&=&0\quad n=2,3,\ldots
\label{dresseden0}
\eea
where
\begin{equation}
R(x)=\int_{-\infty}^\infty\frac{ d\omega}{2\pi} \frac{\exp(i\omega
x)}{1+\exp(U|\omega|/2)}\ .
\end{equation}
The vanishing of the dressed energies of $\Lambda$-strings of lengths
greater than one i.e. $\epsilon_n(\Lambda)=0$ for $n\geq 2$ is due to
the absence of a magnetic field. For finite magnetic fields all
$\epsilon_n(\Lambda)$ will be nontrivial functions.
In order to characterize the excitation spectrum we need to determine
the dressed momenta in addition to the dressed energies. This can be
done by considering the zero temperature limit or \r{densities}
\bea
\rho(k)&=&\frac{1}{2\pi}+\int_{-Q}^Qdk^\prime \cos k\
R(\sin k^\prime -\sin k) \ \rho(k^\prime)\ ,\quad |k|\leq Q\ ;
\qquad \rho(k) = 0 \ , \quad |k| > Q \ ,\nonumber\\
\rho^h(k)&=&\frac{1}{2\pi}+\int_{-Q}^Qdk^\prime \cos k\
R(\sin k^\prime -\sin k) \ \rho(k^\prime)\ ,\quad |k|> Q\ ;
\qquad \rho^h (k) = 0 \ , \quad |k| \le Q \ ,\nonumber\\
\sigma_1(\Lambda)&=&\int_{-Q}^Qdk \frac{1}{U}
\frac{1}{\cosh\frac{2\pi}{U}(\Lambda-\sin k)}\rho(k)\ ,\nonumber\\
\sigma_n(\Lambda)&=&0\quad n=2,3,\ldots\qquad
\sigma_m^h(\Lambda)=0=\sigma_m^\prime(\Lambda)\quad m=1,2,\ldots\qquad\nonumber\\
{\sigma_n^\prime}^h(\Lambda)&=&\frac{1}{\pi}
{\rm Re}\frac{1}{\sqrt{1-(\Lambda -inU/4)^2}}-\int_{-Q}^Qdk\ \frac{nU/4}
{\pi} \frac{\rho(k)}{(nU/4)^2+(\sin k-\Lambda)^2}\ .
\label{densT=0}
\eea
Equations \r{densT=0} describe the ground state of the repulsive
Hubbard model at zero temperature and zero magnetic field. There is
one Fermi sea of $k$'s (charge degrees of freedom) with Fermi rapidity
$\pm Q$ and a second Fermi sea of $\Lambda^1$'s (spin degrees of
freedom), which are filled on the entire real axis.
The total momentum \r{mom} can be rewritten
by using \r{t1}-\r{t3} in the following useful manner
\bea
P&=&\frac{2\pi}{L}\left(
\sum_{j=1}^{N-2M^\prime}I_j+\sum_{m=1}^\infty \sum_{\beta=1}^{M_m}
J^m_\beta
-\sum_{n=1}^\infty \sum_{\alpha=1}^{M_n'} {J^\prime}^n_\alpha
\right)+\pi\sum_{n=1}^\infty \sum_{\alpha=1}^{M'_n}(n+1)\nonumber\\
&=&
\sum_{j=1}^{N-2M^\prime}y(k_j)+\sum_{m=1}^\infty \sum_{\beta=1}^{M_m}
z_m(\Lambda^m_\beta)
-\sum_{n=1}^\infty \sum_{\alpha=1}^{M_n'}
z^\prime_n({\Lambda^\prime}^n_\alpha)
+\pi\sum_{n=1}^\infty \sum_{\alpha=1}^{M'_n}(n+1).
\label{mtmint}
\eea
Using this expression for the total momentum we now can identify the
dressed momenta of various types of excitations. We find that an
additional real $k$ with $|k|>Q$ (``particle'') or hole in the sea of
$k$'s ($|k|<Q$) carry momentum $\pm{\tt p}(k)$ respectively, where
\begin{equation}
{\tt p}(k)=y(k)=2\pi\int_0^k dk^\prime\ [\rho(k^\prime)+\rho^h(k^\prime)]
\ .
\label{pk}
\end{equation}
Similarly, the dressed momentum of a hole in the sea of $\Lambda^1$'s
is
\bea
{\tt p}_1(\Lambda)&=&-z_1(\Lambda)=2\pi\int_\Lambda^\infty
d\Lambda^\prime \sigma_1(\Lambda^\prime)-z_1(\infty)
=2\int_{-Q}^Q {dk}
\arctan[\exp\left(-\frac{2\pi}{U}(\Lambda-\sin k)\right)]\ \rho(k)
-\pi\frac{N}{2L}\ .
\eea
This result was first obtained by Coll \cite{Coll74}.
Finally, a $k$-$\Lambda$ string of length $n$ has dressed momentum
\bea
{\tt p}^\prime_n(\Lambda)&=&-z^\prime_n(\Lambda)+\pi (n+1)\nonumber\\
&=&-2{\rm Re}\arcsin(\Lambda-inU/4)+\int_{-Q}^Qdk\
2\arctan\left(\frac{\Lambda-\sin k}{nU/4}\right)\rho(k)+\pi(n+1),
\label{klmom}
\eea
where the second line is obtained from the $T\to 0$ limit of \r{z3}.
We are now in a position to completely classify the excitation
spectrum at zero temperature. The dispersion relations of all elementary
excitations follow from \r{epsprime}-\r{dresseden0} and \r{pk}-%
\r{klmom}. These equations involve only the two unknown functions,
$\rho (k)$ and $\kappa (k)$, which are solutions of linear Fredholm
integral equations.
\subsection{Ground state for a less than half-filled band}
The integral equations describing the root densities of the ground
state are \r{densT=0}
\bea
\rho(k)&=&\frac{1}{2\pi}+\int_{-Q}^Qdk^\prime \cos k\
R(\sin k^\prime -\sin k) \ \rho(k^\prime)\ ,\quad |k|\leq Q\ ,\nonumber\\
\sigma_1(\Lambda)&=&\int_{-Q}^Qdk \frac{1}{U}
\frac{1}{\cosh\frac{2\pi}{U}(\Lambda-\sin k)}\rho(k)\ ,
\label{densGS}
\eea
where
\begin{equation}
\int_{-Q}^Q dk \rho(k)=N/L\ ,\qquad
\int_{-\infty}^\infty d\Lambda \sigma_1(\Lambda)=M_1/L=N/2L\ .
\end{equation}
The ground state energy per site is given by
\begin{equation}
(E-\mu N)/L=\int_{-Q}^Q dk\ (-2\cos k-\mu -U/2)\ \rho(k)\ .
\end{equation}
\subsection{Excitations for a less than half-filled band}
There are three different kinds of {\sl elementary} excitations.
\begin{itemize}
\item{} The first type of elementary excitation is gapless and
involves the charge degrees of freedom only. It
corresponds to adding a particle to or making a hole in the
distribution of $k$'s.
Such excitations have dressed energy $\mp\kappa(k)$ and dressed
momentum ${\tt p}_{p,h}(k)$.
\item{} The second type of elementary excitation is gapless and carries
spin but no charge. It corresponds to a hole in the distribution of
$\Lambda^1$'s. Such excitations are called {\sl spinons}. They have
dressed energy $-\epsilon_1(\Lambda)$ and dressed momentum ${\tt
p}_1(\Lambda)$.
\item{} There is an infinite number of different types of gapped
excitations that carry charge but no spin. they correspond to adding a
$k$-$\Lambda$ string of length $n$ to the ground state. Their dressed
energies are $\epsilon^\prime_n(\Lambda)$, their dressed momenta are
${\tt p}^\prime_n(\Lambda)$.
\end{itemize}
Let us emphasize that this is a classification of {\sl elementary}
excitations in the repulsive Hubbard model below half-filling.
It is important to distinguish these from ``physical'' excitations,
which are the {\sl permitted combinations} of elementary excitations.
In other words, not any combination of particle-hole excitations,
spinons and $k$-$\Lambda$ strings is allowed, but only those
consistent with the {\sl selection rules} \r{r1}.
To illustrate how this works and relate our findings to known results
in the literature we consider several examples. We introduce the
following terminology: we call the set $\{
N,M_n,M^\prime_n|n=1\ldots\infty\}$ of the numbers of real $k$'s,
$\Lambda$-strings of length $n$ and $k$-$\Lambda$-strings of length
$n$ {\sl occupation numbers} of the corresponding excitation. This is
in contrast to our usage of the term {\sl quantum numbers}, which is
reserved for the eigenvalues of energy, momentum, $S^z, {\vec
S}\cdot{\vec S},\eta^z$ and ${\vec \eta}\cdot{\vec \eta}$.
\vskip .5cm
\underbar{Example 1}: Particle-hole excitation.
This is a two-parametric gapless physical excitation with spin and
charge zero, i.e. its quantum numbers as well as its occupation
numbers are the same as the ones of the
ground state. It is obtained by removing a spectral parameter $k_h$ with
$|k_h|<Q$ from the ground-state distribution of $k$'s and adding a
spectral parameter $k_p$ with $|k_p|>Q$. Its energy and momentum are
\bea
E_{ph}&=&\kappa(k_p)-\kappa(k_h)\ ,\nonumber\\
P_{ph}&=&{\tt p}(k_p)-{\tt p}(k_h)\ .
\eea
This excitation is allowed by the selection rules \r{r1} as in the
ground state only the (half-odd) integers $|I_j|\leq (N-1)/2$ are
occupied and thus the possibility of removing a root corresponding to
$|I_h|\leq (N-1)/2$ and adding a root corresponding to $|I_p|>(N-1)/2$
exists. This excitation was first studied by a different approach in
\cite{Coll74}. In Fig.~\ref{fig:ph} we show the particle-hole spectrum for
densities $n=0.6$ and $n=0.8$. As we approach half-filling the phase-space
for particles shrinks to zero.
\begin{figure}
\begin{center}
\includegraphics[width = 8cm,
height = 8cm ,angle = 0]{ph60.eps}
\includegraphics[width = 8cm,
height = 8cm ,angle = 0]{ph80.eps}
\caption{
Particle-hole excitation for $U=2.0$ and densities $n=0.6$ and
$n=0.8$. Shown are the lower and upper boundaries of the continuum.
}
\label{fig:ph}
\end{center}
\end{figure}
\vskip .5cm
\underbar{Example 2a}: Spin triplet excitation.
Let us consider an excitation involving the spin degrees of freedom
next. If we change the number of down spins by one while keeping the
number of electrons fixed we obtain an excitation with spin
1. Recalling that in the ground state we have $N$ electrons out of
which $M_1=N/2$ have spin down, the excited state will have occupation
numbers $N$ and $M_1=N/2-1$. The selection rules \r{r1} then
read
\begin{equation}
-\frac{L}{2}<I_j\leq \frac{L}{2}\ ,\qquad
|J^1_\alpha|\leq\frac{N}{4}\ .
\end{equation}
The first condition is irrelevant as we are below half filling, but
the second one tells us that there are two more vacancies than there
are roots. In other words, flipping one spin leads to {\sl two} holes
in the distribution of $\Lambda^1$'s. There is one more subtlety we
have to take care of: changing the number of down spins by one, while
keeping the number of electrons fixed leads to a shift of all $I_j$ in
\r{t1} by either $\frac{1}{2}$ or $-\frac{1}{2}$. The consequence of
this shift is a constant contribution of $\pm \pi \frac{N}{L}$ to the
momentum of the excited state. This leads to two ``branches'' of the
same excitation.
The physical excitation obtained in this way is a gapless two-spinon
scattering state with energy and momentum
\bea
E_{\rm trip}&=&-\epsilon_1(\Lambda_1)-\epsilon_1(\Lambda_2)\ ,\nonumber\\
P_{\rm trip}&=&{\tt p}_1(\Lambda_1)+{\tt
p}_1(\Lambda_2)\pm\pi\frac{N}{L}
\ .
\eea
In Fig.~\ref{fig:st} we show the spin-triplet spectrum for
densities $n=0.6$ and $n=0.8$.
\begin{figure}
\begin{center}
\includegraphics[width = 8cm,
height = 8cm ,angle = 0]{st60.eps}
\includegraphics[width = 8cm,
height = 8cm ,angle = 0]{st80.eps}
\caption{
Spin-triplet excitation for $U=2.0$ and densities $n=0.6$ and $n=0.8$.
Shown are the lower and upper boundaries of the continuum for the positive
branch. The negative branch is obtained by reversing the sign of the
momentum. Note that we have not folded back to the first Brillouin zone.
}
\label{fig:st}
\end{center}
\end{figure}
In the Hubbard model the spin-triplet excitations were first studied
by Ovchinnikov \cite{Ovchinnikov70} and by Coll \cite{Coll74}.
The situation encountered here is similar to the spin-1/2 Heisenberg
chain \cite{FaTa84} in the sense that the spin-triplet excitation is a
scattering continuum of two spin-1/2 objects. Furthermore there is a
spin-singlet excitation, which is precisely degenerate with the
triplet (see Example 2b below). This fits nicely into a
picture based on spin-1/2 objects: scattering states of two spinons
give precisely one spin $1$ and one spin $0$ multiplet
$\frac{1}{2}\otimes\frac{1}{2}=1\oplus 0$. Finally, when we
approach half-filling, the spin-triplet continuum constructed above
goes over into the $S=1$ two-spinon scattering continuum of the
half-filled Hubbard model \cite{Woynarovich83a,EsKo94b}.
On the other hand there are differences as well: in the less than
half-filled Hubbard model the Fermi momentum is generally
incommensurate, which leads to incommensurabilities in the spin
excitations (see Fig.~\ref{fig:st}).
More importantly, it is always possible to combine {\sl
any} type of excitation with a particle-hole excitation. It is
therefore not possible to distinguish the two-parametric spin-triplet
excitation constructed above from the special case of a
four-parametric excitation, where a particle-hole excitation sits
``on top'' of the spin-triplet excitation and where the momenta of the
particle and the hole are fixed at the Fermi rapidity.
In other words, due to the presence of gapless particle-hole
excitations there is an inherent ambiguity in the interpretation of the
excitation spectrum on the basis of an ${\cal O}(1)$ calculation of
energy eigenvalues of the Hamiltonian.
\vskip .5cm
\underbar{Example 2b}: Spin singlet excitation.
Let us now choose the occupation numbers as $N$, $M_1=\frac{N}{2}-2$,
$M_2=1$. The corresponding state has the same quantum numbers
as the ground state. From \r{r1} we find that there are $N/2$ vacancies
for real $\Lambda$'s and thus two holes corresponding to rapidities
$\Lambda_1$ and $\Lambda_2$. In other words the excitation considered
involves two spinons. As far as the 2-string is
concerned we find that the associated integer must be zero
$J^2_1=0$. The same shift as in example 2a occurs for the $I_j$'s.
Using \r{dresseden0} and \r{mtmint} we obtain the
energy and momentum of the associated excitation
\bea
E_{\rm sing}&=&-\epsilon_1(\Lambda_1)-\epsilon_1(\Lambda_2)\ ,\nonumber\\
P_{\rm sing}&=&{\tt p}_1(\Lambda_1)+{\tt p}_1(\Lambda_2)\pm\pi\frac{N}{L}\ .
\eea
We see that the spin singlet is precisely degenerate with the
spin triplet considered above. This is a consequence of the
spin $SU(2)$ symmetry of the Hamiltonian in zero magnetic field.
\vskip .5cm
\underbar{Example 3}: $k$-$\Lambda$ string of length $2$.
Let us consider the simplest excitation involving a
$k$-$\Lambda$-string. One possibility is to choose the occupation numbers
as $N$, $M_1=N/2-1$, $M^\prime_1=1$. In addition we keep the distribution
of $I_j$ fixed in such a way that $I_{j+1}-I_j=1$. It is easily
checked that this excitation is allowed by \r{r1}.
Its energy and momentum are
\begin{equation}
E_{k-\Lambda}=\epsilon_1^\prime(\Lambda)\ ,\quad
P_{k-\Lambda}={\tt p}^\prime_1(\Lambda)\ ,
\label{klenmom}
\end{equation}
where $\Lambda\in(-\infty,\infty)$.
\begin{figure}
\begin{center}
\includegraphics[width = 8cm,
height = 8cm ,angle = 0]{kl1.eps}
\includegraphics[width = 8cm,
height = 8cm ,angle = 0]{kl2.eps}
\caption{
Dispersion of a $k$-$\Lambda$ string excitation of length $2$ for
several values of $U$ and density $n$.
}
\label{fig:kl1}
\end{center}
\end{figure}
In Fig.~\ref{fig:kl1} the dispersion of a $k$-$\Lambda$ string of
length 2 is shown for $U=0.5$ and $U=2.0$ and several values of
density $n$. We see that the range of momenta collapses to zero as we
approach half-filling. At the same time the dressed energy approaches
zero. This is in agreement with the results for a half-filled band
\cite{EsKo94b}, where both the dressed energy and the range of
momentum are identically zero.
In order to further exhibit this collapse we subtract the offset
$-2\mu-U$. The resulting curves are displayed in Fig.~\ref{fig:kl3}.
\begin{figure}
\begin{center}
\includegraphics[width = 8cm,
height = 8cm ,angle = 0]{kl3.eps}
\includegraphics[width = 8cm,
height = 8cm ,angle = 0]{kl4.eps}
\caption{
Dispersion of a $k$-$\Lambda$ string excitation of length $2$ for
several values of $U$ and density $n$, where the contribution
$-2\mu-U$ has been subtracted.
}
\label{fig:kl3}
\end{center}
\end{figure}
\subsection{Excitations in the half-filled band}
In the limit of a half-filled band $\mu\to 0$ the Fermi-rapidity $Q$
tends to $\pi$ and the excitation spectrum simplifies drastically
\cite{EsKo94b}. We find that $\epsilon^\prime_n(\Lambda)=0\ \forall n$ and
the only non-vanishing dressed energies are
\bea
\epsilon_1(\Lambda) &=&-
2\int_0^\infty\frac{d\omega}{\omega}\frac{J_1(\omega)\
\cos(\omega\Lambda)}{\cosh(\omega U/4)}\ ,\nonumber\\
\kappa(k) &=& -2\ {\rm cos}k -U/2 - 2 \int_{0}^\infty {d\omega\over \omega}
{J_1(\omega) {\rm cos}(\omega\ {\rm sin}k)e^{-\omega U/4}\over {\rm
cosh}(\omega U/4)}\nonumber\\
\eea
where $J_{0,1}$ are Bessel functions. The charge excitations are now
gapped as a result of the Mott-Hubbard transition.
The {\sl complete} spectrum of physical excitations is derived in
detail in \cite{EsKo94b} (see in particular the appendix).
It is given in terms of spinon and holon scattering states forming
representations of SO(4).
\subsection{Relation to the root-density formalism}
There exists a second standard approach to the zero temperature excitation spectrum
in Bethe Ansatz solvable models, which in the case of the Hubbard model has been
employed in for example in \cite{Ovchinnikov70,Coll74}. We call this the {\sl
root-density formalism} (RDF).
In this approach one specifies a distribution of (half-odd) integers $I_j$, $J^n_\alpha$,
$J'^n_\alpha$ in the discrete Takahashi equations \r{t1}-\r{t3}. One then
takes the thermodynamic limit using that the distribution of roots
becomes dense, so that \r{t1}-\r{t3} turn into a set of coupled linear integral equations
for root densities. From these one calculates energy and momentum via the thermodynamic
limit of \r{mom} and \r{en}.
We will now show how to relate the results of the Yang-Yang approach we implemented above
to the RDF. For definiteness we consider the example of a $k$-$\Lambda$ string excitation
of length $n$.
We start by rewriting the integral equation for $\kappa(k)$ in the following way
\begin{equation}
(1-K)*\kappa\bigg|_k=
\int_{-Q}^Q dk^\prime\left[ \delta(k-k^\prime)-\cos k^\prime\ R(\sin
k^\prime - \sin k)\right] \kappa(k^\prime) = -2 \cos k - \mu -U/2\ .
\label{kappa2}
\end{equation}
Here the kernel of $K$ is given by $K(x,y)=\cos y R(\sin y-\sin x)$.
Equation \r{kappa2} is solved by the Neumann series
\begin{equation}
\kappa(k)=-\sum_{l=0}^\infty \int_{-Q}^Qdk^\prime (2\cos
k^\prime+\mu+U/2) K^l(k,k^\prime)\ ,
\label{neumann}
\end{equation}
where $K^l(x,y)$ denotes the $l$-fold convolution of the kernel
$K(x,y)$. Using \r{neumann} in \r{epsprime} we obtain the following representation
for the dressed energy of a $k$-$\Lambda$ string of length $n$
\bea
&&\epsilon^\prime_n(\Lambda)=
{4{\rm Re}\sqrt{1-(\Lambda -inU/4)^2}-2n\mu-nU}\nonumber\\
&&-\sum_{l=0}^\infty\int_{-Q}^Qdk\int_{-Q}^Qdk^\prime\
\frac{\cos k}{\pi} \frac{nU/4}{(nU/4)^2+(\sin k-\Lambda)^2}
K^l(k,k^\prime)\ (2\cos k^\prime+\mu+U/2)\ .
\label{epsprime2}
\eea
Let us now define a function $\rho^{bs}_n(k|\Lambda)$ by
\begin{equation}
\int_{-Q}^Q dk^\prime\left[ \delta(k-k^\prime)-K(k^\prime,k)\right]
\rho^{bs}_n(k^\prime|\Lambda) = \frac{\cos k}{\pi}
\frac{nU/4}{(nU/4)^2+(\sin k-\Lambda)^2}\ .
\label{rbs}
\end{equation}
The solution of \r{rbs} is given by the Neumann series
\begin{equation}
\rho^{bs}_n(k|\Lambda)=\sum_{l=0}^\infty\int_{-Q}^Q\frac{dk^\prime}{\pi}
K^l(k^\prime,k)\ \cos k^\prime
\frac{nU/4}{(nU/4)^2+(\Lambda-\sin k^\prime)^2}\ .
\label{rhobs}
\end{equation}
Using \r{rhobs} in \r{epsprime2} we finally obtain
\begin{equation}
\epsilon^\prime_n(\Lambda)=
{4{\rm Re}\sqrt{1-(\Lambda -inU/4)^2}-2n\mu-nU}
-\int_{-Q}^Qdk (2\cos k+\mu+U/2)\ \rho^{bs}_n(k|\Lambda)\ .
\end{equation}
\subsection{Summary}
In this section we have reviewed the Yang-Yang approach \cite{YaYa69}
to the thermodynamics of the Hubbard model \cite{Takahashi72}. We have
shown how to express the Gibbs free energy per site \r{gibbs} in terms
of the solutions of the infinite set of coupled nonlinear integral
equations \r{zeta}-\r{kl}.
By taking the zero-temperature limit of the thermodynamic equations
we have obtained the complete classification of the spectrum of
elementary excitations of the Hubbard model below half-filling for
vanishing magnetic field.
We emphasize that this approach is based on Takahashi's string
hypothesis \cite{Takahashi72} (see section II.C). It does not only
provide the dispersion curves of all elementary excitations in the
zero temperature limit, but also a set of `selection rules'
\r{t1}-\r{t3}. These rules determine the set of {\sl physical}
excitations, i.e.\ the allowed combinations of elementary excitations
(see section VI.E and, for the half-filled case, \cite{EsKo94a,%
EsKo94b}).
The numerical calculation of the Gibbs free energy from \r{gibbs} or
\r{gibbs2} is difficult, since it involves the solution of an infinite
number of coupled non-linear integral equations \r{zeta}-\r{kl}.
A truncation scheme has to be introduced \cite{KUO89,UKO90}, which
restricts the numerical accuracy of the calculation.
In the next section we present a different approach to the
thermodynamics of the Hubbard model, which circumvents such
difficulties.
\section{Thermodynamics in the quantum transfer approach}
In this section we present a treatment of the thermodynamic
properties of the Hubbard chain in a lattice path integral
formulation with a subsequent eigenvalue analysis of the matrix
describing transfer along the chain direction
(quantum transfer matrix, QTM). This approach has several advantages.
First, the analysis is very simple as only the largest
eigenvalue of the QTM is necessary in order to calculate the free
energy. This has to be compared with the traditional TBA
\cite{Takahashi72}
(see Sect.\ VI) where all eigenvalues of the Hamiltonian have to be
taken into account. Second, the number of `density functions' and
integral equations obtained below is {\it finite} in contrast to
\cite{Takahashi72}
(see Sect.\ VI A). Finally, the finite temperature correlation lengths
can be derived through a calculation of next-largest eigenvalues of the
QTM
\cite{Klu92,Klu93}. This however, will not be explored in this report.
Within the QTM approach we obtain several results of which we want
to point out the conceptual achievements. First, the data obtained
for various physical quantities agree with those obtained in
\cite{KUO89,UKO90} based on Takahashi's string hypothesis. Judging from
this we do not see any evidence for a failure of Takahashis's
formulation based on strings. Second, in the low temperature limit our
integral equations quite naturally yield the Tomonaga-Luttinger liquid
picture of separate spin and charge contributions to the free energy as
suggested in \cite{FrKo90,FrKo91}. Mathematically, the dressed
energy formalism known from the ground state analysis is recovered.
\subsection{The classical counterpart}
There are many direct path integral formulations of the Hubbard model,
see for instance \cite{Shastry86a,Bariev82}. For our purposes we want
to keep the integrability structure as far as possible. To this end it
proved useful to employ the exactly solvable classical model
corresponding to the Hubbard chain, namely Shastry's model which is a
two-sublattice six-vertex model with decoration \cite{Shastry88b}.
More precisely, the vertex weights of the classical model for the case
$U=0$ are given by the product of the vertex weights of two six-vertex
models (with components $\sigma$ and $\tau$): $\ell_{1,2} (u) =
\ell^{\sigma}_{1,2}(u) \otimes \ell^{\tau}_{1,2}(u)$, where
\begin{equation}
\ell^{\sigma}_{1,2}(u) = \tst{\2} (\cos(u) + \sin(u)) +
\tst{\2} (\cos(u) - \sin(u)) \sigma^z_1 \sigma^z_2 +
(\sigma_1^+ \sigma_2^- + \sigma_1^- \sigma_2^+) \qd.
\end{equation}
Taking account of the $U\not=0$ interactions, the following local
vertex weight operator (denoted by $S$) was found \cite{Shastry88b}
\begin{eqnarray}
S_{1,2}(v,u)
&=& \cos(u+v)\, \ch(h(v,U)-h(u,U))\, \ell_{1,2} (v-u)
+ \cos(v-u)\, \sh(h(v,U)-h(u,U))\, \ell_{1,2} (u+v)\,
\sigma_2^z \tau_2^z ,
\end{eqnarray}
where $\sh(2 h(u,U)) := \frac{U}{4} \sin(2u)$.
The Yang-Baxter equation for triple $S$ matrices
was conjectured \cite{Shastry88b}, but only recently proved in
\cite{ShWa95}.
The so-called $L$-operator is related to $S$ by
\begin{equation} \label{shastryl}
L_{i,g}(u)=S_{i,g}(u,0) \qd,
\end{equation}
where $i$ and $g$ are indices referring to the $i$th lattice site and
the auxiliary space, respectively. For the $L$-operator the proof of
the Yang-Baxter equation with $S$ as intertwiner
was already given in \cite{Shastry88b}.
The commutativity of the row-to-row transfer matrix
\begin{equation} \label{shastryt}
{\cal T}(u) := \tr\prod^{\leftarrow}_i L_{i, g} (u)
\end{equation}
is a direct consequence.
Next we define $R_{1,2}(u,v) = \Pi_{1,2} S_{1,2}(v,u)|_{U \rightarrow
- U}$, where $\Pi_{1,2}$ is the permutation matrix. The matrix elements
$R_{\alpha,\beta}^{\mu,\nu}(u,v)$ will be considered as the local
Boltzmann weights associated with vertex configurations $\alpha$,
$\beta$, $\mu$, $\nu$ on the lower, upper, left, and right bond, where
the spectral parameters $u$ and $v$ ``live'' on the vertical and
horizontal bonds, respectively.
For later use we introduce $\overline{R}(u,v)$ and $\widetilde{R}(u,v)$
($u$ and $v$ associated with the vertical and horizontal bond)
by clockwise and anticlockwise 90$^\circ$ rotations of $R$,
or in matrix notation
\begin{equation}
\overline{R}_{\alpha,\beta}^{\mu,\nu}(u,v)=
R_{\mu,\nu}^{\beta,\alpha}(v,u) \qd, \qd
\widetilde{R}_{\alpha,\beta}^{\mu,\nu}(u,v)=
R_{\nu,\mu}^{\alpha,\beta}(v,u) \qd.
\end{equation}
Similar to (\ref{shastryl}) and (\ref{shastryt}) we can associate a
row-to-row transfer matrix with $\overline{R}$. We note the Hamiltonian
limits ${\cal T}(u)=\exp(iP+uH+O(u^2))$ and $\overline{\cal T}(u) =
\exp(-iP+uH+O(u^2))$. Consequentially, the partition function of the
Hubbard chain at finite temperature $T = 1/ \beta$ is given by
\begin{equation}
Z = \lim_{L\rightarrow \infty} \tr \, e^{-\beta H} =
\lim_{L\rightarrow \infty}\lim_{N\rightarrow \infty}
\hbox{ tr } [{\cal T}(u) \overline{\cal T}(u)]^{N/2} |_{u=\beta/N}
\qd.
\end{equation}
We regard the resulting system as a fictitious two-dimensional model
on a $L\times N$ square lattice. Here $N$ is the extension in the
fictitious (imaginary time) direction, sometimes referred to as the
Trotter number. The lattice consists of alternating rows each being a
product of only $R$ weights or of only $\overline{R}$ weights,
respectively.
Now by looking at the system in a 90$^\circ$ rotated frame which turns
$\overline{R}$ and $R$ weights into $R$ and $\widetilde{R}$ weights,
it is natural to define
the `quantum transfer matrix' (QTM) by
\begin{equation}
\{ {\cal T}_{{\rm QTM}} (u,v) \}^{\{\beta\}}_{\{\alpha\}}
:= \sum_{\{ \mu \}} \prod_{i=1}^{N/2}
R^{\mu_{2i-1},\mu_{2i}}_{\alpha_{2i-1},\beta_{2i-1}} (-u,v) \,
\widetilde{R}^{\mu_{2i},\mu_{2i+1}}_{\alpha_{2i},\beta_{2i}}(u,v)
\qd,
\label{QTM}
\end{equation}
which is identical to the column-to-column transfer matrix of the
square lattice for $v=0$. The interchangeability of the two limits
($ L, N \rightarrow \infty$) \cite{SuIn87,SAW90} leads to the
following expression for the partition function,
\begin{equation}
Z= \lim_{N\rightarrow \infty} \lim_{L\rightarrow \infty}
\; \tr \left[{\cal T}_{{\rm QTM}}\left(u=\frac{\beta}{N},0\right)
\right]^{L} \qd.
\end{equation}
There is a gap between the largest and the second largest eigenvalues
of ${\cal T}_{{\rm QTM}}(u,0)$ for finite $\beta$. Therefore the free
energy per site is expressed just by the largest eigenvalue
$\Lambda_{{\rm max}}(u,0)$ of ${\cal T}_{{\rm QTM}}(u,0)$,
\begin{equation} \label{freeen}
f = - \, T \lim_{N\rightarrow \infty}\ln\Lambda_{{\rm max}}
\left(u=\frac{\beta}{N},0\right) \qd.
\end{equation}
It is relatively simple to see that (\ref{QTM}) is integrable, i.e.\ a
family of commuting operators for variable $v$ and fixed $u$.
A non-vanishing chemical potential $\mu$ and magnetic field $B$
can be incorporated\footnote{We note that our conventions for the
magnetic field in this article are different from \cite{JKS98}. In
\cite{JKS98} the magnetic field was denoted by $H = 2B$.}
as they merely lead to trivial modifications due to twisted boundary
conditions for the QTM (cf.\ \cite{Klu93}).
\subsection{Diagonalization of the Quantum Transfer Matrix}
Here we summarize the main results of \cite{JKS98} where the
diagonalization of (\ref{QTM}) on the basis of an algebraic Bethe ansatz
was performed. Note that the general expression for the eigenvalue
of the quantum transfer matrix is quite complicated \cite{JKS98}, but
simplifies considerably at $v=0$ and $u \rightarrow 0$,
\begin{equation}
\Lambda(v=0) = e^{\beta U/4}
(1+e^{\beta (\mu+B)} )(1+e^{\beta (\mu-B)} )
u^N \prod_{j=1}^{m} z_j \qd.
\label{eig_prac}
\end{equation}
The numbers $z_j$ are charge rapidities satisfying Bethe ansatz
equations which are most transparently written in terms of the related
quantities
\begin{equation}
s_j =\frac{1}{2i} \left(z_j-\frac{1}{z_j}\right) \qd.
\end{equation}
For these rapidities $s_j$ and additional rapidities $w_{\alpha}$ the
coupled eigenvalue equations read
\begin{equation}
e^{-\beta(\mu - B)} \phi(s_j) =
-\frac{q_2(s_j - \i U/4)}{q_2(s_j + \i U/4)} \qd, \qd
e^{- 2\beta\mu} \,
\frac{q_2(w_\alpha + \i U/2)}{q_2(w_\alpha - \i U/2)} =
- \frac{q_1(w_\alpha + \i U/4)}{q_1(w_\alpha - \i U/4)} \qd,
\label{QTMBA}
\end{equation}
where we have employed the abbreviations for products over rapidities
\beq
q_1(s)=\prod_j(s-s_j) \qd, \qd q_2(s)=\prod_\alpha(s-w_\alpha) \qd.
\eeq
The function $\phi$ is defined by
\begin{equation}
\phi(s) =
\left( \frac{(1-z_-/z(s))(1-z_+/z(s))}
{(1+z_-/z(s))(1+z_+/z(s))} \right)^{N/2} \qd, \qd
z(s) = \i s \left( 1+\sqrt{(1-1/s^2)} \right) \qd,
\end{equation}
where $z(s)$ possesses two branches. The standard (``first'') branch
is chosen by the requirement $z(s)\simeq 2\i s$ for large values of
$s$, and the branch cut line $[-1,1]$.
The numbers $z_\pm$ are defined by $z_\pm=\exp(\alpha)(\tan u)^{\pm 1}$,
where $\sinh(\alpha)= -{U\over 4}\sin 2u$.
As usual, equations like (\ref{QTMBA}) which are identical in structure
to (\ref{bak}) and (\ref{bas}) have many solutions. In our approach to
the thermodynamics just the largest eigenvalue of the QTM matters. The
corresponding distribution of rapidities is relatively simple. For
$\mu=B=0$ the rapidities are all situated on the real axis. Naturally,
for finite $\mu$ and $B$ we have modifications which, however, do not
affect qualitative aspects of the distribution.
In a similar way a path integral formulation of the Hubbard chain
and a diagonalization of the corresponding QTM was employed in
\cite{Tsune91,KlBa96}. In \cite{Tsune91} the eigenvalue equations
were studied numerically for finite Trotter number and the case of
half-filling. In \cite{KlBa96} an analytic attempt was undertaken to
study the limit of infinite Trotter number and to derive a set of
non-linear integral equations. Unfortunately, these equations were
rather ill posed with respect to numerical evaluations. In the next
section we present a formulation of non-linear integral equations on
the basis of recent work \cite{JKS98}.
\subsection{Non-linear integral equations}
\label{sec:non-linear-integral-equations}
In this section we are concerned with the derivation of well posed
integral equations equivalent to (\ref{QTMBA}) for the largest
eigenvalue of the QTM and thus for the free energy per site (see
(\ref{freeen})). We introduce a set of auxiliary functions (${\frak b}$, ${\frak c}$,
and $\overline{{\frak c}}$) described in more detail in (\ref{auxFunct}) below. These
auxiliary functions are complex functions, however mostly evaluated
close to the real axis.
In terms of the auxiliary functions the Gibbs free energy per site is
expressed in several ways
\bea
f & = & - \mu - \frac{U}{4} - \frac{T}{2 \p \i}
\int_{\cal L}[\ln z(s)]' \ln\left(1+{\frak c}+\overline{{\frak c}}\right)ds \nonumber \\
& & \qqqd + \, \frac{T}{4 \p \i} \int_{\cal L}
\left[\ln\frac{z(s-\i U/2)}{z(s)}\right]'\ln(1+{\frak b}(s))ds
+ \frac{T}{4 \p \i} \int_{\cal L}
\left[\ln\frac{z(s+\i U/2)}{z(s)}\right]'\ln(1+1/{\frak b}(s))ds
\qd, \label{int-eig-p} \\
& = & \frac{U}{4} - \frac{T}{2 \p \i} \int_{\cal L}[\ln z(s)]'
\ln\frac{1+{\frak c}+\overline{{\frak c}}}{\overline{{\frak c}}}ds
- \frac{T}{2 \p \i}
\int_{\cal L} \left[\ln{z(s-\i U/2)}\right]'\ln(1+{\frak c}(s))ds
\qd.
\label{int-eig}
\eea
For yet another expression see \cite{JKS98}.
Equations (\ref{int-eig-p}) and (\ref{int-eig}) have to be compared
with equations (\ref{gibbs}) and (\ref{gibbs2}) which give the free
energy in the string based approach of Takahashi. In contrast to the
string based approach, the auxiliary functions ${\frak b}$, ${\frak c}$ and $\overline{{\frak c}}$
entering (\ref{int-eig-p}) and (\ref{int-eig}) satisfy a closed set of
finitely many (non-linear) integral equations,
\bea
\ln{\frak b}&=&-2 \beta B+K_2\squareforqed\ln(1+{\frak b})-{\overline{K_1}}\circ
\ln(1+1/\overline{{\frak c}}) \qd, \nonumber \\
\ln{\frak c}&=&-\beta{U}/{2}+\beta(\mu+B)+\varphi
-{\overline{K_1}}\squareforqed\ln(1+1/{\frak b})-{\overline{K_1}}\circ\ln(1+\overline{{\frak c}}) \qd,
\nonumber \\
\ln\overline{{\frak c}}&=&-\beta{U}/{2}-\beta(\mu+B)-\varphi
+K_1\squareforqed\ln(1+{\frak b})+K_1\circ\ln(1+{\frak c}) \qd.
\label{all}
\eea
Here we have used the definition
\begin{equation}
\varphi(x)=-2\beta \i x\sqrt{1-1/x^2}
\end{equation}
and have introduced the notation
\begin{equation}
(g\circ f)(s)=\int_{\cal L}g(s-t)f(t)dt
\end{equation}
for the convolution of two functions $g$ and $f$
with contour ${\cal L}$ surrounding the real axis
at infinitesimal distance above and below in anticlockwise
manner. The definition of $\squareforqed$ is similar, with integration
contour surrounding the real axis at imaginary parts $\pm U/4$.
The kernel functions are rational functions,
\beq
K_1(s)=\frac{U/4 \pi}{s(s + \i U/2)} \qd, \qd
{\overline{K_1}}(s)=\frac{U/4 \pi}{s(s - \i U/2)} \qd, \qd
K_2(s)=\frac{U/2 \pi}{s^2 + U^2/4} \qd.
\label{kernels}
\eeq
Next, we want to point out that the function ${\frak b}$ will be evaluated
on the lines $\Im s=\pm U/4$. The functions ${\frak c}$ and $\overline{{\frak c}}$ need only
be evaluated on the real axis infinitesimally above and below the
interval $[-1,1]$. Also the convolutions involving the ``${\frak c}$
functions'' in (\ref{all}) can be restricted to a contour surrounding
$[-1,1]$ as these functions are analytic outside.
Lastly, we want to comment on the derivation of (\ref{int-eig}),
(\ref{all}).
The explicit expressions of the functions ${\frak b}$, ${\frak c}$, $\overline{{\frak c}}$
are
\bea
{\frak b}&=&
\frac{\bl{1}+\bl{2}+\bl{3}+\bl{4}}{l_1+l_2+l_3+l_4} \qd, \nonumber
\\
{\frak c}&=&\frac{l_1+l_2}{l_3+l_4}\cdot
\frac{\bl{1}+\bl{2}+\bl{3}+\bl{4}}
{l_1+l_2+l_3+l_4+\bl{1}+\bl{2}+\bl{3}+\bl{4}} \qd,
\label{auxFunct} \\
\overline{{\frak c}}&=&\frac{\bl{3}+\bl{4}}{\bl{1}+\bl{2}}\cdot
\frac{l_{1}+l_{2}+l_{3}+l_{4}}
{l_1+l_2+l_3+l_4+\bl{1}+\bl{2}+\bl{3}+\bl{4}} \qd, \nonumber
\eea
where the functions $l_j$ and $\bl{j}$ are
\beq
l_j(s)=\lambda_j(s- \i U/4)\cdot{\rm e}^{2 \beta B}{\phi^+(s)\phi^-(s)} \qd,
\qd \bl{j}(s)=\lambda_j(s+ \i U/4) \qd,
\eeq
and the $\lambda_j$ are defined in terms of the $q_1$ and $q_2$
functions, i.e. in terms of the Bethe ansatz rapidities
\bea
\lambda_1(s) &=&{\rm e}^{\beta(\mu+B)}
\frac{\phi(s- \i U/4)}{q_1(s- \i U/4)} \qd, \qd
\lambda_2(s) ={\rm e}^{2\beta\mu}
\frac{q_2(s- \i U/2)}{q_2(s)q_1(s- \i U/4)} \qd, \nonumber \\
\lambda_3(s) &=&\frac{q_2(s+ \i U/2)}{q_2(s)q_1(s+ \i U/4)} \qd,
\qquad\quad
\lambda_4(s) ={\rm e}^{\beta(\mu-B)}
\frac{1}{\phi(s+ \i U/4)\,q_1(s+ \i U/4)} \qd.
\label{eig-aux}
\eea
The functions defined in (\ref{auxFunct}) are proven to
satisfy a set of closed functional
equations which can be transformed into integral form (\ref{all}),
cf.\ \cite{JKS98}.
Also (\ref{int-eig}) follows from (\ref{auxFunct})
after a lengthy yet direct
calculation. The merit of (\ref{int-eig}),
(\ref{all}) is that this formulation does no longer make any reference
to the Bethe ansatz equations! Hence the calculation of an infinite
set of discrete rapidities is replaced by the computation of analytic
functions for which much more powerful tools are available.
\subsection{Analytical solutions of the integral equations}
\label{sec:analytical-solutions}
Before numerically studying
various thermodynamic properties for general temperatures
and particle concentrations we want to give some analytic treatments
of limiting cases of the Hubbard model.
\subsubsection{Strong-coupling limit}
\label{sec:strong-coupling-limit}
In the strong-coupling limit \mbox{$U\to\infty$} at half-filling
($\mu=0$) the Hubbard model is expected to reduce to the
Heisenberg chain. Indeed, in the strong-coupling limit
we find that ${\frak c}$, $\overline{{\frak c}}\to 0$. Hence, the only non-trivial
function determining the eigenvalue of the QTM is ${\frak b}$,
see the first expression of (\ref{int-eig}). The integral equation
for ${\frak b}$ as obtained from (\ref{all}) is identical to
that obtained directly for the thermodynamics of the Heisenberg model
\cite{JKS98}.
\subsubsection{Free-Fermion limit}
\label{sec:free-fermion-limit}
Next, let us consider the limit $U\to{}0$ leading to a free fermion
model, however representing a non-trivial consistency check of the
equations. Indeed, the auxiliary functions can be calculated
explicitly. Finally, the free energy per site reads
\beq
f = - \frac{T}{2\pi} \int_{-\pi}^{\pi}
\ln \Big\{ \big[1 + \exp((\mu + B + 2\cos{k})/T) \big]
\big[1 + \exp((\mu - B + 2\cos{k})/T) \big] \Big\}
\, {\rm d}k \qd,
\eeq
which, as desired, is the result for free tight binding electrons.
\subsubsection{Low-temperature asymptotics}
\label{sec:low-temperature-limit}
The low-temperature regime is the most interesting limit as the system
shows Tomonaga-Luttinger liquid behavior. We want to describe the
relation of the non-linear integral equations to the known dressed
energy formalism \cite{FrKo90,FrKo91} of the Hubbard model. This
represents a further and in fact the most interesting consistency check.
For $T=1/\beta\to 0$ we can simplify the non-linear integral equations
as they turn into {\it linear} integral equations, however in the
generic case ($B$ and $\mu\not=0$) with {\it finite} integration
contours. This can be seen as follows. To be specific let us adopt
fields $B>0$, $\mu\le 0$ (particle density $n\le 1$). In the
low-temperature limit several auxiliary functions tend to zero, namely
${{\frak b}(s)}\to{}0$ on the line $\Im\, s=-U/4$, ${\frak c}(s)\to 0$ above and
below the real axis, and $1/\overline{{\frak c}}(s)\to{}0$ just below the real axis,
i.e.\ on the line $\Im\, s=-\epsilon$. The remaining non-trivial
functions are ${{\frak b}}$ on the line $\Im\, s=+U/4$ and $1/\overline{{\frak c}}$ just above
the real axis for which we introduce the notation
\beq
b(\lambda)={\frak b}(\lambda+ \i U/4) \qd, \qd
c(\lambda)=1/\overline{{\frak c}}(\lambda+ \i \epsilon) \qd.
\eeq
We note that
\beq
|b|,\, |c|\gg{}1
\quad{\rm for}\quad |x|<\lambda_0,\, {\sigma}_0
\quad{\rm and}\quad
|b|,\, |c|\ll{}1\quad{\rm for}\quad |x|>\lambda_0,\, {\sigma}_0 \qd,
\eeq
for certain crossover values $\lambda_0,\, {\sigma}_0$. The slopes for
the crossover are steep, so that the following approximations to the
integral equations (\ref{all}) are valid,
\bea \label{linearbc-p}
\ln b & = & \phi_b - \int_{-\lambda_0}^{\lambda_0}
k_2 (\lambda - \lambda')\, \ln b (\lambda')\,
{\rm d}\lambda'
+ \int_{-k_0}^{k_0} k_1 (\lambda - \sin k')\, \cos k' \,
\ln c(k')\, {\rm d}k' \qd, \\
\ln c & = & \phi_c + \int_{-\lambda_0}^{\lambda_0}
k_1 (\sin k - \lambda')\, \ln b (\lambda')\,
{\rm d}\lambda' \qd.
\label{linearbc}
\eea
where $k_1(\lambda)=K_1(\lambda- \i U/4)=\overline{K_1}(\lambda+
\i U/4)$ and $k_2(\lambda)=K_2(\lambda)$. In order to facilitate
comparison with the dressed energy formalism we also introduced
a new integration variable $k$ leading to a change of the boundaries
of integration, $\s_0 \rightarrow k_0 = \arcsin \s_0$. The driving
terms in (\ref{linearbc-p}) and (\ref{linearbc}) are related to the
bare energies
\beq
\varepsilon_s^0 = B \qd, \qd
\varepsilon_c^0 = - 2 \cos k - \m - U/2 - B
\eeq
by
\begin{equation}
\phi_b =- \beta\,\varepsilon_s^0+ \CO(1/\beta) \quad{\rm and}\quad
\phi_c =- \beta\,\varepsilon_c^0+ \CO(1/\beta) \qd.
\end{equation}
Therefore, we find the following connections between auxiliary
functions and the dressed energy functions,
\begin{equation}
\ln b = - \beta\,\varepsilon_s + \CO(1/\beta) \qd, \qd
\ln c = - \beta\,\varepsilon_c + \CO(1/\beta) \qd.
\label{BC-relations-in-low-T-limit}
\end{equation}
For a comparison with \cite{FrKo90,FrKo91} note the different
normalization of the chemical potential.
For a comparison with the results in section VI we set the magnetic
field equal to zero. Then $\varepsilon_s^0 = B = 0$, and it can be seen
that $\lambda_0 = \infty$. Setting $Q = k_0$ we can identify
(\ref{linearbc-p}), (\ref{linearbc}) in the zero temperature limit
with (\ref{dresseden0}). We find
\bea
&& - \lim_{T \rightarrow 0} T \ln b (\la) =
\lim_{T \rightarrow 0} \epsilon_1 (\la) =
\lim_{T \rightarrow 0} T \ln \h_1 (\la) \qd, \\
&& - \lim_{T \rightarrow 0} T \ln c (\la) =
\lim_{T \rightarrow 0} \k (k) =
\lim_{T \rightarrow 0} T \ln \z (k) \qd.
\eea
The free energy also admits the above approximation scheme, yielding
up to $\CO(T^2)$-terms the low-temperature expansion
\begin{equation}
f=\varepsilon_0 -\frac{\pi}{6}\,
\left(\frac{1}{v_c}+\frac{1}{v_s}\right)T^2 \qd.
\label{lowTsep}
\end{equation}
Here the definitions of the sound velocities and the ground state
energy are standard. The additive occurrence of $1/v_c$ and $1/v_s$
on the right hand side of (\ref{lowTsep}) is a manifestation of
spin-charge separation in the one-dimensional Hubbard model, due to
which each elementary excitation contributes independently to
(\ref{lowTsep}). The velocities $v_c$ and $v_s$ typically take
different values.
\subsubsection{High-temperature limit}
Finally, we consider the high-temperature limit $T\to\infty$ with $B$,
$U$ as well as $\beta\mu$ fixed ensuring a fixed particle density $n$.
The integral equations turn into algebraic equations which are easily
solved, resulting in the high-temperature limit
\begin{equation}
S = 2 \ln \left( \frac{2}{2-n} \right) -
n \ln \left( \frac{n}{2-n} \right)
\end{equation}
for the entropy, as expected by counting the degrees of freedom per
lattice site. Especially at half-filling, $n=1$, this equals to
$S=\ln(4)$.
\subsection{Numerical Results}
Here we show numerical results for
specific heat $C$, magnetic susceptibility
$\chi_m$, and charge susceptibility $\chi_c$
for half-filling and small doping,
see Figs.\ \ref{fig:fig1},\ \ref{fig:fig2},\ \ref{fig:fig3}.
In addition, we aim at a comparison of results
\cite{KUO89,UKO90} obtained within the thermodynamic Bethe ansatz
\cite{Takahashi72} based on the string hypothesis, and results
\cite{JKS98} obtained within the quantum transfer matrix approach.
In essence, we observe convincing agreement of the data obtained
within the two absolutely different approaches. We conclude that
there is no indication
for any failure of Takahashi's
formulation of thermodynamics based on the string hypothesis.
Of course rather small differences exist in the sets of data. A more
thorough comparison shows deviations of the data presented in
Fig.\ \ref{fig:fig1} and\ \ref{fig:fig2}.
This is nothing but expected due to the truncation procedure adopted
in \cite{KUO89,UKO90}. Instead of dealing with an infinite
set of integral equations for
density functions (\ref{densities},\ref{zeta}-\ref{kl})
for string excitations of spin rapidities
(\ref{ideal1})
and charge rapidities
(\ref{idealk}), a finite subset was taken into account. In the case
of complex charge rapidities, strings describe excitations with gap.
These degrees of freedom are less sensitive to errors introduced by the
truncation than strings involving only
spin rapidities which describe gapless excitations. In fact, the
agreement of the data for the charge susceptibility $\chi_c$
is best, small deviations are observed in $C$ and
$\chi_m$.
The advantage of the QTM approach over the traditional TBA is threefold.
First, in \cite{JKS98} the Hubbard chain with very strong doping
was analyzed (apparently not possible in the traditional TBA) and quite
unexpected structures in the susceptibility data were found. In the
magnetic susceptibility only traces of spinon excitations were visible.
The charge susceptibility, however, exposed holon signatures and
additional maxima due to spinon excitations indicating deviations from
the concept of spin-charge separation. Second, the accuracy of
numerical data obtained within the QTM approach is much higher as it
is more efficient to deal with a set of integral equations which is
strictly finite from beginning. Finally, even correlation lengths can
be calculated within the QTM approach although not presented in detail
yet.
\newlength{\Breite}
\setlength{\Breite}{5cm}
\begin{figure}
\begin{picture}(0,200)
\put(13.0,16.0){\includegraphics[width=1.461\Breite,
height=1.265\Breite,angle=0]{spec2.eps}}
\put(220,0){
\includegraphics[width=1.5\Breite,
height=1.5\Breite,angle=0]{c2.8.eps}}
\end{picture}
\caption{Specific heat $C$ versus temperature for $U = 8$ and particle
densities $n=1$, $n=0.9$, $n=0.8$ and $n=0.7$. The left graph shows
results obtained within the string based approach,
the right graph shows results obtained by the quantum transfer
matrix method.
}
\label{fig:fig1}
\end{figure}
\begin{figure}
\begin{picture}(0,200)
\put(1,11){
\includegraphics[width=1.529\Breite,
height=1.32\Breite,angle=0.5]{magsusc2.eps}}
\put(220,0){
\includegraphics[width=1.5\Breite,
height=1.5\Breite,angle=0]{chi2.8.eps}}
\end{picture}
\caption{Magnetic susceptibility $\chi_m$ versus temperature for $U = 8$
and particle densities $n=1$, $n=0.9$, \dots, $n=0.5$. The left graph
shows results obtained within the string based approach,
the right graph shows results obtained by the quantum transfer
matrix method.
}
\label{fig:fig2}
\end{figure}
\begin{figure}
\begin{picture}(0,200)
\put(-2.5,7){
\includegraphics[width=1.532\Breite,
height=1.336\Breite,angle=0]{chargsus2.eps}}
\put(220,0){
\includegraphics[width=1.5\Breite,
height=1.5\Breite,angle=0]{kap2.8.eps}}
\end{picture}
\caption{Charge susceptibility $\chi_c$ versus temperature for
$U = 8$ and particle densities $n=1$, $n=0.9$, $n=0.8$ and $n=0.7$.
The left graph shows results obtained within the string based approach,
the right graph shows results obtained by the quantum transfer matrix
method.
}
\label{fig:fig3}
\end{figure}
\subsection{Equivalence of string based thermodynamics with QTM
approach}
Here we want to address the fundamental and perhaps puzzling question
how two exact, however completely different formulations of the
thermodynamics of the Hubbard model as presented in sections~VI and VII
can exist.
It is very important to understand this problem as the
means of derivation and the mathematical properties of both approaches
are seemingly lacking any similarities.
In the remainder of this section we restrict ourselves to a sketch
of the general mathematical structures underlying
the relation of the traditional thermodynamical formulation based on
typically infinitely many density functions and the new formulation
with strictly finitely many auxiliary functions. This relation
has been worked out for a number of models including spin chains
\cite{Klu92} and correlated fermion models like $t$-$J$ systems
\cite{JKS98b}, not yet however for the Hubbard chain. In the following
we want to introduce the general technique for spin-1/2 Heisenberg
models and related systems.
The starting point is an integrable Hamiltonian corresponding to an
exactly solvable classical model as discussed in section~VII.A. For
this model a path integral representation is formulated leading to an
integrable QTM {\boldmath $T$}$(v)$ with eigenvalue equation
\begin{equation}
T(v)={f(v-\lambda)q(v+\lambda)+f(v)q(v-\lambda)\over q(v)} \qd,
\label{Tq}
\end{equation}
where $f(v)$ is a known function and $q(v)$ is a polynomial or a
product over trigonometric functions with zeros exactly at the
Bethe ansatz rapidities. By a line of reasoning similar to that
presented in VII.C {\it one} non-linear integral equation can be
derived which yields full information on the free energy of the model.
The key to understanding the relation of the standard TBA to the above
approach is the notion of fusion algebras and inversion identities.
By repeated use of the fusion procedure a set of transfer matrices
{\boldmath $T$}$(v)=$ {\boldmath $T$}$^1(v)$,
{\boldmath $T$}$^2(v)$, {\boldmath $T$}$^3(v)$,...,
is generated satisfying
\begin{equation}
\mbox{\boldmath $T$}^{q}(v)\mbox{\boldmath $T$}^{q}(v+\lambda)=
f(v-\lambda) f(v+\lambda)\mbox{\boldmath $I$}
+\mbox{\boldmath $T$}^{q+1}(v)\mbox{\boldmath
$T$}^{q-1}(v+\lambda) \qd, \label{func}
\end{equation}
where $q$ takes integer values, $q = 1, 2, \dots$, and
{\boldmath $T$}$^0(v)$ is proportional to the identity operator.
Introducing operators
\begin{equation}
\mbox{\boldmath $Y$}^{q}(v)={\mbox{\boldmath $T$}^{q-1}(v+\lambda)
\mbox{\boldmath $T$}^{q+1}(v)\over
f(v-\lambda) f(v+q\lambda)}\label{littlet}
\end{equation}
we find the extremely useful inversion identity hierarchy
\begin{equation}
\mbox{\boldmath $Y$}^{q}(v)\mbox{\boldmath $Y$}^{q}(v+\lambda)
=[\mbox{\boldmath $I$}+\mbox{\boldmath $Y$}^{q-1}(v+\lambda)]
[\mbox{\boldmath $I$}+\mbox{\boldmath $Y$}^{q+1}(v)] \qd.
\label{invIdHier}
\end{equation}
This set of functional equations can be transformed into a set of
non-linear integral equations for the largest eigenvalue of
{\boldmath $T(v)$}. These integral equations appear to be identical to
the equations obtained along the traditional TBA approach!
Several remarks are in order. Relations (\ref{invIdHier}) were derived
in \cite{Zamolodchikov91a,Zamolodchikov91b} starting with the TBA
equations, however the connection to the fusion hierarchies was not
made. In contrast to the TBA integral equations the functional
equations (\ref{invIdHier}) (as of course equation (\ref{Tq})) admit
more than one solution. The physical meaning of these solutions
becomes clear only through the microscopic construction of transfer
matrices. The next-leading eigenvalues describe the correlation
lengths of static correlation functions at finite temperature. Such
applications are investigated in current research. Some results have
been reported in \cite{JKS98b,FKM98,SSSU99,Sakai99}.
In summary, the traditional TBA equations are not in contradiction with
the new thermodynamics based on the QTM. Rather on the contrary, the
QTM approach represents a unified approach to both sets of integral
equations. The central object is the study of the largest eigenvalue
of the QTM. This can be achieved by two different methods, either by
a Bethe ansatz (\ref{Tq}) or by use of the fusion hierarchy
(\ref{func}).
In physical terms, the appearance of the fusion hierarchy is related
to the pole structure of the intertwiner, i.e. the $R$-matrix satisfying
the YBE. In turn the $R$-matrix is related to the $S$-matrix of
particles, the poles are related to bound states (strings) which in
turn are the central objects of the traditional analysis of the
thermodynamics of integrable systems. In this way we observe
mathematical and physical connections between the initially so
differently looking approaches.
\section{Derivation of the wave function}
\subsection{General setting}
The expression (\ref{sectorq})-(\ref{wwfsa}) for the Bethe ansatz wave
function of the Hubbard model was first presented by Woynarovich
\cite{Woynarovich82a}. The purpose of this appendix is to give a
detailed derivation of Woynarovich's wave function. We will make use
of results obtained in references \cite{LiWu68}, \cite{Yang67} and of
the quantum inverse scattering method \cite{KBIBo,EKS92a}. Other
derivations of the Bethe Ansatz wave functions for the Hubbard model
can be found for example in \cite{Sutherland85} and in \cite{IzSk:88}.
The outline of this appendix is as follows.
In section 1 we define the wave function $\psi(x_1,\ldots, x_N)$,
for $N$ electrons and derive the first quantized Schr\"odinger
equation for the Hubbard model.
In sections $2$ and $3$ we present the explicit solution of the
Schr\"odinger equation with periodic boundary conditions for the cases
of $2$ and $3$ electrons, respectively. We treat these cases in
considerable detail for pedagogical reasons. Finally, in section $4$
we discuss the general case of $N$ electrons.
An important ingredient in the construction of the wave function is
the exact solution of an inhomogeneous spin-1/2 Heisenberg model. The
essence of the nested Bethe Ansatz procedure employed in constructing
wave-functions for the Hubbard Hamiltonian is the reduction of this
problem to a simpler one, which involves only the spin degrees of
freedom. The dynamics of the spin degrees of freedom are described by
an inhomogeneous Heisenberg model, and its exact solution constitutes
the ``nesting'' of the Bethe Ansatz procedure. We summarize the
algebraic Bethe Ansatz solution of the inhomogeneous Heisenberg model
in Appendix \ref{section:XXX}.
Let us recall the explicit form of the Hamiltonian
\beq \label{hamilapp}
H = - \sum_{j=1}^L \sum_{\s = \auf, \ab}
(c_{j, \s}^+ c_{j+1, \s} + c_{j+1, \s}^+ c_{j, \s})
+ U \sum_{j=1}^L
(n_{j \auf} - \tst{\2})(n_{j \ab} - \tst{\2})\ .
\eeq
As the total number of electrons $N$ and the number of electrons with
spin down $M$ are good quantum numbers, we can use them to label
eigenstates of \r{hamilapp}
\bea
|N,M \rangle = \sum_{\{\sigma_j\}} \sum_{\{x_k\}}
\psi(x_1\ldots x_N; \sigma_1,\ldots,\sigma_N)
c_{x_1, \sigma_1}^{\dagger} \cdots c_{x_N, \sigma_N}^{\dagger} |0
\rangle\ .
\label{eigenstate}
\eea
Here $\sum_{\{\sigma_j\}}$ denotes summation over all $N!/((N-M)!M!)$
possible spin-configurations with $M$ down-spins.
Due to the anticommutation relations between the Fermion operators,
we may assume without loss of generality that the amplitudes $\psi$
are totally antisymmetric
\beq
\psi(x_{P_1},\ldots,x_{P_N} ; \sigma_{P_1},\ldots,\sigma_{P_N})
= \sign(P) \psi(x_{1},\ldots,x_{N} ; \sigma_{1},\ldots,\sigma_{N})
\ ,
\label{antisymmetry}
\eeq
where $P=(P_1,P_2,\ldots, P_N)$ is a permutation of the labels
$\{1,2,\ldots, N\}$, i.e. an element of the symmetric group $S_N$.
The antisymmetry property \r{antisymmetry} implies that the summation
over spin configurations in \r{eigenstate} is redundant. Indeed one
finds that
\beq
|N,M \rangle = {\frac {N!} {(N-M)!M!}} \sum_{\{x_j\}}
\psi(x_{1},\ldots,x_{N} ; \sigma_{1},\ldots,\sigma_{N})
c_{x_1, \sigma_1}^{\dagger} \cdots
c_{x_N, \sigma_N}^{\dagger} |0 \rangle\ ,
\eeq
where $(\sigma_1,\ldots,\sigma_N)\in S_N$ is arbitrary. In order to
derive the Schr\"odinger equations it is therefore convenient to work
with the following simplified expression for general eigenstates of $H$
\beq
|N,M;{\vec \sigma} \rangle = \sum_{\{ x_j\}}
\psi(x_{1},\ldots,x_{N} ; \sigma_{1},\ldots,\sigma_{N})
c_{x_1, \sigma_1}^{\dagger} \cdots c_{x_N, \sigma_N}^{\dagger} |0
\rangle \ .
\eeq
It is now straightforward to show, that the eigenvalue problem
\beq
H |N,M;{\vec \sigma}\rangle = E |N,M;{\vec \sigma}\rangle\ ,
\eeq
implies the following Schr\"odinger equation for the wave function
$\psi$ \cite{LiWu68}
\bea
&& - \sum_{j=1}^{N} \sum_{s=\pm 1}
\psi(x_1, \ldots, x_j + s, \ldots, x_N; {\vec \sigma} ) +
U \sum_{j<k} \delta(x_j, x_k)
\psi(x_{1},\ldots,x_{N} ; \sigma_{1},\ldots,\sigma_{N})
\nonumber\\
&&\qquad = (E+\frac{UN}{2}-\frac{UL}{4}) \psi(x_{1},\ldots,x_{N};
\sigma_{1},\ldots,\sigma_{N}) \ .
\label{SG}
\eea
Here $\delta(a,b)$ denotes the Kronecker delta.
\subsection{\boldmath Two electrons}
Let us now explicitly construct the wave function for the case of
two electrons, $N=2$. The Schr\"odinger equation is of the form
\bea
&&-\psi(x_1-1, x_2; \sigma_1,\sigma_2 )
-\psi(x_1+1, x_2; \sigma_1,\sigma_2 )
-\psi(x_1, x_2-1; \sigma_1,\sigma_2 )
-\psi(x_1, x_2+1; \sigma_1,\sigma_2 )\nonumber\\
&&\qquad +U \delta(x_1, x_2) \psi(x_1,x_2;\sigma_1,\sigma_2)
= (E+U-\frac{UL}{4}) \psi(x_1,x_2;\sigma_1,\sigma_2)\quad .
\label{2SG}
\eea
As long as $x_1<x_2$ or $x_1>x_2$ \r{2SG} reduces to the Schr\"odinger
equation for free electrons on a lattice and its solutions are
therefore just superpositions of plane waves. When the electrons
occupy the same site, they interact. This can be thought of in terms
of a scattering process. Due to integrability this scattering is
purely elastic, which means that the momenta of the two electrons are
individually conserved. Thus, the most that can happen is that the
electrons exchange their momenta. These considerations lead to the
famous ``nested'' Bethe ansatz form for the wave functions, which we
will discuss next.
Let $Q$ be a permutation of the labels of coordinates i.e.
$Q=(Q_1,Q_2)\in\{ (1,2), (2,1)\}$. In the ``sector'' $Q$ defined by
the condition $x_{Q_1}\leq x_{Q_2}$ the nested Bethe Ansatz for the
wave function is
\beq
\psi(x_1, x_2; \sigma_1,\sigma_2) =
\sum_{P \in S_2}\sign(PQ) A_{\sigma_{Q_1}\sigma_{Q_2}}(k_{P_1},k_{P_2})
\exp(i \sum_{j=1}^{2} k_{Pj}x_{Qj}) \quad .
\label{BAWF2}
\eeq
Substituting \r{BAWF2} into \r{2SG} for the case $x_1\neq x_2$ we obtain
\beq
E = - ( 2 \cos k_1 + 2 \cos k_2) -U+\frac{UL}{4}\qd.
\eeq
When $x_1=x_2$ we have to ``match'' the wave function defined in the
two sectors $Q=(12)$ and $Q=(21)$. This requires single valuedness
\bea
\psi(x,x;\sigma_1,\sigma_2)&=&
\left[A_{\sigma_1\sigma_2}(k_1,k_2)-A_{\sigma_1\sigma_2}(k_2,k_1)\right]
\exp(i[k_1+k_2]x)\nonumber\\
&=& \left[A_{\sigma_2\sigma_1}(k_2,k_1)-A_{\sigma_2\sigma_1}(k_1,k_2)\right]
\exp(i[k_1+k_2]x)\ .
\label{cont}
\eea
In addition, the Schr\"odinger equation \r{2SG} for $x=x_1=x_2$ needs
to be fulfilled, which yields the condition
\bea
&&-e^{-ik_1}A_{\sigma_1\sigma_2}(k_1,k_2)
+e^{-ik_2}A_{\sigma_1\sigma_2}(k_2,k_1)
+e^{ik_2}A_{\sigma_2\sigma_1}(k_1,k_2)
-e^{ik_1}A_{\sigma_2\sigma_1}(k_2,k_1)\nonumber\\
&&-e^{ik_2}A_{\sigma_1\sigma_2}(k_1,k_2)
+e^{ik_1}A_{\sigma_1\sigma_2}(k_2,k_1)
+e^{-ik_1}A_{\sigma_2\sigma_1}(k_1,k_2)
-e^{-ik_2}A_{\sigma_2\sigma_1}(k_2,k_1)\nonumber\\
&&+U\left[A_{\sigma_1\sigma_2}(k_1,k_2) -
A_{\sigma_1\sigma_2}(k_2,k_1)\right]\nonumber\\
&&= -2(\cos k_1+\cos k_2)\left[A_{\sigma_1\sigma_2}(k_1,k_2) -
A_{\sigma_1\sigma_2}(k_2,k_1)\right].
\label{2SG2}
\eea
By means of \r{2SG2} and \r{cont} we can express two of the four
amplitudes $A_{\sigma_{Q_1}\sigma_{Q_2}}(k_{P_1},k_{P_2})$ in terms of
the other two. A short calculation gives
\bea
A_{\sigma_1\sigma_2}(k_2,k_1)&=&
\frac{-U/2i}{\sin k_1 - \sin k_2 -U/2i} A_{\sigma_1\sigma_2}(k_1,k_2)
+\frac{\sin k_1 - \sin k_2}{\sin k_1 - \sin k_2 -U/2i}
A_{\sigma_2\sigma_1}(k_1,k_2)\ .
\label{ybe0}
\eea
Equation \r{ybe0} has a natural interpretation in terms of a
scattering process of two particles. In order to see this we rewrite
it as
\bea
A_{\sigma_2\sigma_1}(k_2,k_1)&=&
\sum_{\tau_1,\tau_2} S^{\sigma_1\tau_1}_{\sigma_2\tau_2}(k_1,k_2)\
A_{\tau_1\tau_2}(k_1,k_2)\ ,
\label{ybe1}
\eea
where $S(k_1,k_2)$ is the two-particle S-matrix with elements
\begin{equation}
S^{\sigma_1\tau_1}_{\sigma_2\tau_2}(k_1,k_2)=
\frac{-U/2i}{\sin k_1 - \sin k_2 -U/2i}
\Pi^{\sigma_1\tau_1}_{\sigma_2\tau_2}
+\frac{\sin k_1 - \sin k_2}{\sin k_1 - \sin k_2 -U/2i}
I^{\sigma_1\tau_1}_{\sigma_2\tau_2}\ .
\label{smat}
\end{equation}
Here $I$ is the identity operator
$I^{\sigma_1\tau_1}_{\sigma_2\tau_2}=\delta_{\sigma_1,\tau_1}
\delta_{\sigma_2,\tau_2}$ and $\Pi$ is a permutation operator
$\Pi^{\sigma_1\tau_1}_{\sigma_2\tau_2}=\delta_{\sigma_1,\tau_2}
\delta_{\sigma_2,\tau_1}$.
In the next step we want to impose periodic boundary conditions on the
wave function
\bea
\psi(L+1,x_2;\sigma_1,\sigma_2)&=&\psi(1,x_2;\sigma_1,\sigma_2)\ ,\nonumber\\
\psi(0,x_2;\sigma_1,\sigma_2)&=&\psi(L,x_2;\sigma_1,\sigma_2)\ ,\nonumber\\
\psi(x_1,L+1;\sigma_1,\sigma_2)&=&\psi(x_1,1;\sigma_1,\sigma_2)\ ,\nonumber\\
\psi(x_1,0;\sigma_1,\sigma_2)&=&\psi(x_1,L;\sigma_1,\sigma_2)\ .
\eea
A short calculation shows that this imposes the following conditions on
the amplitudes
\begin{equation}
A_{\sigma_{Q_1}\sigma_{Q_2}}(k_{P_1},k_{P_2})=\exp\left(ik_{P_1}L\right)
A_{\sigma_{Q_2}\sigma_{Q_1}}(k_{P_2},k_{P_1})\ ,
\label{pbc2A}
\end{equation}
where $P,Q\in S_2$ are arbitrary.
In principle we now could simply solve \r{pbc2A} and thus determine
the quantization conditions for the momenta $k_{1,2}$. Rather than
proceeding in this way, we will introduce some seemingly unnecessary
formalism, which however will be very useful for treating the case of
more than two electrons.
We define an auxiliary spin model on a lattice with $N$ sites i.e. a
lattice formed by the electrons. On every site $j$ there are two
allowed configurations $|\uparrow\rangle_j$ and $|\downarrow\rangle_j$,
corresponding to spin up and down respectively.
Next we define spin operators $S_j^{\pm,z}$ acting on the resulting
Hilbert space as follows
\bea
S_j^{-}|\downarrow\rangle_j =0=S_j^{+}|\uparrow\rangle_j\ ,\quad
S_j^{-}|\uparrow\rangle_j =|\downarrow\rangle_j\ ,\quad
S_j^{+}|\downarrow\rangle_j =|\uparrow\rangle_j\ ,\quad
S_j^{z}|\downarrow\rangle_j =-\frac{1}{2}|\downarrow\rangle_j\ ,\quad
S_j^{z}|\uparrow\rangle_j =\frac{1}{2}|\uparrow\rangle_j\ .
\label{spinops}
\eea
We now define a particular set of states in the spin model in the
following way \cite{EKS92a}
\beq
|k_{P_1},\ldots,k_{P_N} \rangle =
\sum_{\sigma_1\ldots\sigma_N=\pm 1}A_{\sigma_1\ldots \sigma_N}
(k_{P_1},\ldots,k_{P_N})
\prod_{\ell =1 } ^{N} \left(S_{\ell}^{-}
\right)^{(1- \sigma_{\ell})/2} |0 \rangle \qd.
\label{Astates}
\eeq
Here we have used conventions where $\sigma_j=\uparrow$ corresponds to $1$
and $\sigma_j=\downarrow$ corresponds to $-1$. The set of equations
\r{ybe1}, due to the Schr\"odinger equation, now induces the following
between states in the spin model
\beq
|k_{P_1},k_{P_2} \rangle =
Y^{1,2}(\sin k_{P_2}, \sin k_{P_1})
|k_{P_2},k_{P_1} \rangle \quad ,
\label{ybey}
\eeq
where the operator $Y^{j,k}$ is given by
\begin{equation}
Y^{j,k}(v_1, v_2)
= {\frac {-U/2i} {v_1 -v_2 - U/2i}} I
+ {\frac {v_1-v_2} {v_1 -v_2 - U/2i}}
\Pi^{(j,k)} \quad.
\label{yjk}
\end{equation}
Here $I$ is the identity operator over the Hilbert space of the spin
model and $\Pi^{(j,k)}$ is a permutation operator
\beq
\Pi^{(j,k)} = {\frac 1 2}
\left( I + 4 {\vec S}_j \cdot {\vec S}_k \right) \qd.
\label{pi12}
\eeq
Here $S_k^{x,y,z}$ are the spin operators defined in \r{spinops},
where $S^\pm_k=S^x_k\pm i S^y_k$.
In order to derive \r{ybey} one simply uses the explicit form \r{smat}
of the S-matrix and the following properties of permutation operators
\begin{equation}
\Pi^{(1,2)}|0\rangle=|0\rangle\ ,\quad
\Pi^{(1,2)}S^-_1\Pi^{(1,2)}=S^-_2\ ,\quad
\Pi^{(1,2)}S^-_2\Pi^{(1,2)}=S^-_1\ .
\end{equation}
The $Y$-operators (which are related to the S-matrix \r{smat} via
multiplication by a permutation operator) were first introduced by
C.N. Yang in his seminal 1967 paper \cite{Yang67}.
On the level of the auxiliary spin model the periodic boundary
conditions \r{pbc2A} translate into
\begin{equation}
|k_{P_1},k_{P_2}\rangle=\exp\left(ik_{P_1}L\right) \Pi^{(12)}
|k_{P_2},k_{P_1}\rangle\ .
\end{equation}
With the use of \r{ybey} this is then transformed into
\begin{equation}
|k_{P_1},k_{P_2}\rangle=\exp\left(ik_{P_1}L\right)
X^{1,2}(\sin k_{P_1},\sin k_{P_2}) |k_{P_1},k_{P_2}\rangle\ ,
\label{pbc2}
\end{equation}
where the operator $X^{j,k}$ is given by
\bea
X^{j,k}(v_1, v_2)& =& \Pi^{(j,k)}Y^{j,k}(v_1, v_2)
= {\frac {-U/2i} {v_1 -v_2 - U/2i}}
\Pi^{(j,k)}+
{\frac {v_1-v_2} {v_1 -v_2 - U/2i}} I \qd.
\label{xjk}
\eea
In other words, $X^{1,2}$ is the S-matrix \r{smat} viewed as an
operator in the auxiliary spin model.
We now make the crucial observation that the operator $X^{1,2}$ is
precisely the transfer matrix of an inhomogeneous spin-1/2
Heisenberg model on a 2-site lattice
\begin{equation}
X^{1,2}(\sin k_{P_1},\sin k_{P_2})=
\tau(\sin k_{P_1}|\{\sin k_{P_1},\sin k_{P_2}\})\ ,
\end{equation}
where $\tau$ is given by \r{transferm} with $N=2$. This is shown in
full generality in \r{xtau}.
The periodic boundary conditions \r{pbc2A} can thus be
rewritten as an eigenvalue problem for the transfer matrix
$\tau(\Lambda=\sin k_{P_1}|\{\sin k_{P_1},\sin k_{P_2} \})$ of an
inhomogeneous Heisenberg model on 2-site lattice
\beq
|k_{P_1},k_{P_2}\rangle =e^{i Lk_{P_1}}\
\tau(\sin k_{P_1}|\{\sin k_{P_1},\sin k_{P_2} \})\
|k_{P_1},k_{P_2} \rangle \qd.
\label{pbc22}
\eeq
The diagonalization of the transfer matrix $\tau(\Lambda|\{\sin
k_{P_1},\sin k_{P_2} \})$ is carried out in Appendix
\ref{section:XXX}. {From} the point of view of constructing
eigenstates of the Hubbard Hamiltonian with periodic boundary
conditions, we have succeeded in reducing the problem to a much
simpler one, namely diagonalizing the transfer matrix of an
inhomogeneous Heisenberg model. This is the essence of C.N. Yang's
nested Bethe Ansatz procedure.
For the problem at hand we need to distinguish two cases, depending on
the spins of the two electrons.
\begin{itemize}
\item{} \underline{two electrons with spin up}
Here the appropriate eigenstate of $\tau(\Lambda|\{\sin k_{P_1},\sin
k_{P_2} \})$ is found in the sector with no overturned spins, i.e. it
is the ferromagnetic state with all spins up. From \r{eigen} we see
that its eigenvalue is equal to $1$. The periodic boundary conditions
for the case of two electrons with spin up thus take the simple form
\begin{equation}
e^{ik_j L}=1\ ,\qquad j=1,2.
\end{equation}
These are precisely the Lieb-Wu equations \r{bak},\r{bas} for the case
$N=2$, $M=0$.
Using that the appropriate eigenstate of $\tau(\Lambda|\{\sin
k_{P_1},\sin k_{P_2} \})$ is equal to $|0\rangle$,
we infer by comparison with \r{Astates} that all amplitudes are equal
to $1$. This then implies the following explicit form for the wave
function in sector $Q$
\beq
\psi(x_1, x_2; \sigma_1,\sigma_2) =
\sum_{P \in S_2}\sign(PQ) \exp(i \sum_{j=1}^{2} k_{Pj}x_{Qj}) \quad .
\eeq
This agrees with Woynarovich's result \r{wswf}.
\item{} \underline{one electron with spin up, one with spin down}
Now the appropriate eigenstate of $\tau$ is found in the sector with
one overturned spin. Using \r{eigen} we obtain the corresponding
eigenvalue
\begin{equation}
1-iU/[2(\sin k_{P_1}-\lambda_1+iU/4)] = \frac{\sin k_{P_1}-\lambda_1-iU/4}
{\sin k_{P_1}-\lambda_1+iU/4}\ ,
\label{a31}
\end{equation}
where $\lambda_1$ fulfills the Bethe Ansatz equations of the
inhomogeneous Heisenberg model on a 2-site lattice
\bea
1 & = & \prod_{j=1}^{2} {\frac
{\lambda_1 - \sin k_j + iU/4}
{\lambda_1 - \sin k_j - iU/4}} \quad .
\label{lw21}
\eea
Inserting \r{a31} into \r{pbc22} we obtain the following
quantization conditions for the momenta $k_j$ due to periodic boundary
conditions
\bea
\exp(i L k_{P_1}) & = & {\frac {\lambda_1 - \sin k_{P_1} - iU/4}
{\lambda_1 - \sin k_{P_1} + iU/4} } \quad , \qd {\rm for} \, P \in S_2
\quad .
\label{lw22}
\eea
Equations \r{lw21} and \r{lw22} coincide with the Lieb-Wu equations
\r{bak},\r{bas} for the case $N=2$ and $M=1$.
Let us also determine an explicit expression for the amplitudes
$A_{\sigma_1\sigma_2}(k_1,k_2)$. Comparing the result \r{1spin} for
the eigenstate of $\tau(\sin k_{P_1}|\{\sin k_{P_1},\sin k_{P_2} \})$
with \r{Astates} we see that
\begin{equation}
A_{\sigma_1\sigma_2}(k_{P_1},k_{P_2})=
\prod_{j=1}^{y-1} \left({\lambda_1 - \sin k_{P_j} - {iU\over
4}\over {\lambda_1 - \sin k_{P_j} + {iU\over 4}}}\right){1\over\lambda_1 -
\sin k_{P_y} + {iU\over 4}}\ ,
\label{aexpl}
\end{equation}
where $y$ is the position of the down spin in the sequence
$\sigma_1\sigma_2$.
The wave functions \r{BAWF2} with amplitudes \r{aexpl} coincide with
Woynarovich's result \r{wswf} for the case $N=2$, $M=1$.
\end{itemize}
\subsection{\boldmath Three electrons}
Let us now explicitly construct the wave function for the case of
three electrons $N=3$. The Schr\"odinger equation is of the form
\bea
&&- \sum_{s=\pm 1}\left[\psi(x_1+s, x_2, x_3; \sigma_1,\sigma_2,\sigma_3 )
+\psi(x_1, x_2+s, x_3; \sigma_1,\sigma_2,\sigma_3 )
+\psi(x_1, x_2, x_3+s; \sigma_1,\sigma_2,\sigma_3 )\right] \nonumber\\
&&+U \sum_{j<k} \delta(x_j, x_k) \psi(x_1,x_2,x_3; \sigma_1,\sigma_2,\sigma_3)
= (E+\frac{3U}{2}-\frac{UL}{4}) \psi(x_1,x_2,x_3;
\sigma_1,\sigma_2,\sigma_3)\quad .
\label{3SG}
\eea
In the sector defined by $x_{Q_1}\leq x_{Q_2}\leq x_{Q_3}$, where $Q$
is a permutation of the labels $\{1,2,3\}$, the Bethe Ansatz
wavefunction reads
\beq
\psi(x_1, x_2,x_3; \sigma_1,\sigma_2,\sigma_3) =
\sum_{P \in S_3}\sign(PQ)
A_{\sigma_{Q_1}\sigma_{Q_2}\sigma_{Q_3}}(k_{P_1},k_{P_2},k_{P_3})
\exp(i \sum_{j=1}^{3} k_{Pj}x_{Qj}) \quad .
\label{BAWF3}
\eeq
Substituting \r{BAWF3} into \r{3SG} for the case $x_1\neq x_2\neq
x_3\neq x_1$ we obtain the following expression for the energies
\beq
E = - ( 2 \cos k_1 + 2 \cos k_2 +2\cos k_3) -\frac{3U}{2}+\frac{UL}{4}\qd.
\eeq
Single-valuedness of the wave-function now leads to a larger number of
relations between the amplitudes. We have to consider the three cases
$\psi(x,x,x_3;\sigma_1,\sigma_2,\sigma_3)$,
$\psi(x,x_2,x;\sigma_1,\sigma_2,\sigma_3)$,
$\psi(x_1,x,x;\sigma_1,\sigma_2,\sigma_3)$.
A simple calculation yields the following conditions on the amplitudes
\begin{equation}
A_{\sigma_{Q_1}\sigma_{Q_2}\sigma_{Q_3}}(k_{P_1},k_{P_2},k_{P_3})
-A_{\sigma_{Q_1}\sigma_{Q_2}\sigma_{Q_3}}(k_{P'_1},k_{P'_2},k_{P'_3})
=
A_{\sigma_{Q'_1}\sigma_{Q'_2}\sigma_{Q'_3}}(k_{P'_1},k_{P'_2},k_{P'_3})
-A_{\sigma_{Q'_1}\sigma_{Q'_2}\sigma_{Q'_3}}(k_{P_1},k_{P_2},k_{P_3}),
\label{cont3}
\end{equation}
where $Q$ and $P$ are arbitrary permutations of $\{1,2,3\}$ and
$Q'=Q(j,j+1)$, $P'=P(j,j+1)$. Our notations are such that for any
permutation of $N$ elements $S=(S_1,\ldots ,S_N)$
\begin{equation}
S(j,j+1)=(S_1,\ldots, S_{j-1},S_{j+1},S_j,S_{j+2},\ldots,S_N).
\end{equation}
Let us consider a specific example of \r{cont3} in more detail. The
wave-function $\psi(x,x,x_3;\sigma_1,\sigma_2,\sigma_3)$ with $x_3> x$
can be expressed alternatively in sector $Q=(1,2,3)$ or in sector
$Q'=(2,1,3)$. Equating the respective expressions \r{BAWF3} and using
the orthogonality property of plane waves we obtain
\begin{equation}
A_{\sigma_{1}\sigma_{2}\sigma_{3}}(k_{P_1},k_{P_2},k_{P_3})
-A_{\sigma_{1}\sigma_{2}\sigma_{3}}(k_{P_2},k_{P_1},k_{P_3})
=
A_{\sigma_{Q_2}\sigma_{Q_1}\sigma_{Q_3}}(k_{P_2},k_{P_1},k_{P_3})
-A_{\sigma_{Q_2}\sigma_{Q_1}\sigma_{Q_3}}(k_{P_1},k_{P_2},k_{P_3}).
\end{equation}
This indeed coincides with the general result \r{cont3}.
We note that the Bethe ansatz wave function \r{BAWF3} by construction
is antisymmetric under simultaneous exchange of spin and space
variables. It is easy to see that this fact assures the Schr\"odinger
equation \r{3SG} to be satisfied, when the three electron are occupying
the same site. Moreover, this reasoning readily generalizes to the case
of an arbitrary number of electrons in the following subsection. The
only non-trivial case left to consider is the case of two electrons at
the same site.
Let us start with $x_1=x_2=x<x_3$. Using \r{BAWF3} in \r{3SG} we obtain
the following condition on the amplitudes
\bea
&&-[e^{-ik_{P_1}}+e^{ik_{P_2}}]A_{\sigma_1\sigma_2\sigma_3}(k_{P_1},k_{P_2},k_{P_3})
+[e^{-ik_{P_2}}+e^{ik_{P_1}}]A_{\sigma_1\sigma_2\sigma_3}(k_{P_2},k_{P_1},k_{P_3})\nonumber\\
&&+[e^{ik_{P_2}}+e^{-ik_{P_1}}]A_{\sigma_2\sigma_1\sigma_3}(k_{P_1},k_{P_2},k_{P_3})
-[e^{ik_{P_1}}+e^{-ik_{P_2}}]A_{\sigma_2\sigma_1\sigma_3}(k_{P_2},k_{P_1},k_{P_3})\nonumber\\
&&+U\left[A_{\sigma_1\sigma_2\sigma_3}(k_{P_1},k_{P_2},k_{P_3}) -
A_{\sigma_1\sigma_2\sigma_3}(k_{P_2},k_{P_1},k_{P_3})\right]\nonumber\\
&&= -2(\cos k_{P_1}+\cos k_{P_2})\left[A_{\sigma_1\sigma_2\sigma_3}(k_{P_1},k_{P_2},k_{P_3}) -
A_{\sigma_1\sigma_2\sigma_3}(k_{P_2},k_{P_1},k_{P_3})\right].
\label{3sg}
\eea
Now making use of the fact that \r{3sg} and \r{cont3} are of the same
structure as \r{2SG2} and \r{cont}, we conclude that the following
relation between amplitudes holds
\bea
A_{\sigma_2\sigma_1\sigma_3}(k_{P_2},k_{P_1},k_{P_3})&=&
\sum_{\tau_1,\tau_2} S^{\sigma_1\tau_1}_{\sigma_2\tau_2}(k_{P_1},k_{P_2})\
A_{\tau_1\tau_2\sigma_3}(k_{P_1},k_{P_2},k_{P_3})\ ,
\eea
where $S$ is the two-particle S-matrix \r{smat}. All other cases of
coinciding coordinates can be analysed in exactly the same way.
The final result is
\bea
A_{\sigma_{Q'_1}\sigma_{Q'_2}\sigma_{Q'_3}}(k_{P'_1},k_{P'_2},k_{P'_3})&=&
\sum_{\tau_1,\tau_2} S^{\sigma_{Q_j}\tau_1}_{\sigma_{Q_{j+1}}\tau_2}(k_{P_j},k_{P_{j+1}})\
A_{\sigma_{Q_1}\ldots \sigma_{Q_{j-1}}\tau_1\tau_2
\sigma_{Q_{j+2}}\ldots\sigma_{Q_3}}(k_{P_1},k_{P_2},k_{P_3})\ ,
\label{3asa}
\eea
where $Q$ and $P$ are arbitrary permutations and
$Q'=Q(j,j+1)$,$P'=P(j,j+1)$. Equation \r{3asa} has important
consequences. In order to exhibit these clearly, it is convenient to
express \r{3asa} in the framework of the auxiliary spin model
introduced above. Inserting \r{3asa} into \r{Astates} we obtain
\beq
|k_{P'_1},k_{P'_2},k_{P'_3} \rangle =
Y^{j,j+1}(\sin k_{P_{j}}, \sin k_{P_{j+1}})
|k_{P_1},k_{P_2},k_{P_3} \rangle \quad ,
\label{3ybey}
\eeq
where $Y^{j,k}$ is given by \r{yjk}. There are altogether
$(N-1)N!=12$ equations \r{3ybey} and all of them have to be consistent
with one another! This puts severe constraints on the operators
$Y^{j,k}$.
Let us explain this in more detail. We recall that the symmetric
group $S_N$ is generated by the identity $\id$ and the transpositions
of nearest neigbours $(j, j + 1)$, $j = 1, \dots, N - 1$, modulo the
relations
\bea \label{braid}
(j, j + 1) (j + 1, j + 2) (j, j + 1) & = &
(j + 1, j + 2) (j, j + 1) (j + 1, j + 2) \qd, \\ \label{nnncom}
(j, j + 1) (k, k + 1) & = & (k, k + 1) (j, j + 1) \qqd \mbox{for}
\ |j - k| > 1 \qd, \\
(j, j + 1) (j, j + 1) & = & \id \qd. \label{idem}
\eea
Equation (\ref{braid}) is called the braid relation. Note that
(\ref{nnncom}) is non-trivial only for $N > 3$.
By inspection of (\ref{3ybey}) we see that the $Y$-operators act on
the states $|k_1, k_2, k_3\>$ by exchanging neighbouring components of
the vector $\vec k = (k_1, k_2, k_3)$,
that determines the state $|k_1, k_2, k_3\>$ of our auxiliary spin
system. Hence all states $|k_{P_1}, k_{P_2}, k_{P_3}\>$, $P \in S_3$,
can be obtained from $|k_1, k_2, k_3\>$ by repeated use of
(\ref{3ybey}). Equivalently, (\ref{3ybey}) allows us to obtain the
state corresponding to any permutation $\bar P \in S_3$ from a state
corresponding to any other permutation $P \in S_3$.
It follows from (\ref{braid})-(\ref{idem}) that a representation of a
permutation as a product of transpositions of nearest neighbours is
not unique.
For the case at hand, $N = 3$, we have for instance, $(1,3) =
(1,2) (2,3) (1,2) = (2,3) (1,2) (2,3)$. We are thus facing a consistency
problem for equations (\ref{3ybey}): The relations (\ref{braid})-%
(\ref{idem}) impose consistency conditions on the $Y$-operators.
Let us study these consistency conditions. For $N = 3$ we only have to
consider (\ref{braid}) for the case $j=1$ and (\ref{idem}). Thus
(\ref{braid}) implies that
\bea
|k_{\bar{P}_1},k_{\bar{P}_2},k_{\bar{P}_3} \rangle &=&
Y^{1,2}(\sin k_{P_{2}}, \sin k_{P_{3}})
Y^{2,3}(\sin k_{P_{1}}, \sin k_{P_{3}})
Y^{1,2}(\sin k_{P_{1}}, \sin k_{P_{2}})
|k_{P_1},k_{P_2},k_{P_3} \rangle \ ,\nonumber\\
&=&
Y^{2,3}(\sin k_{P_{1}}, \sin k_{P_{2}})
Y^{1,2}(\sin k_{P_{1}}, \sin k_{P_{3}})
Y^{2,3}(\sin k_{P_{2}}, \sin k_{P_{3}})
|k_{P_1},k_{P_2},k_{P_3} \rangle ,
\label{yyy}
\eea
where $\bar{P}= P (1,3) = P (1,2) (2,3) (1,2) = P (2,3) (1,2) (2,3) =
(P_3,P_2,P_1)$. Assuming $|k_1, k_2, k_3\>$ to be arbitrary, we conclude
that
\begin{equation}
Y^{1,2}(s_2,s_3)
Y^{2,3}(s_1,s_3)
Y^{1,2}(s_1,s_2)
=
Y^{2,3}(s_1,s_2)
Y^{1,2}(s_1,s_3)
Y^{2,3}(s_2,s_3)\ ,
\label{yyy2}
\end{equation}
where $s_{1,2,3}$ are arbitrary complex numbers. This is the famous
Yang-Baxter equation. It is now crucial that (\ref{yyy2}) can be
verified by direct calculation. Therefore (\ref{3ybey}) is consistent
with (\ref{braid}).
Similarly, the use of equation (\ref{idem}) in (\ref{3ybey}) yields
the requirement
\begin{equation}
\left(Y^{j,j+1}(u,v)\right)^{- 1}=Y^{j,j+1}(v,u)\ .
\label{yinv}
\end{equation}
It is easy to verify by direct calculation that the operators $Y^{j,k}$
defined in\r{yjk} indeed fulfil \r{yinv}. Hence (\ref{3ybey}) is
consistent with (\ref{idem}), and we conclude that the entire set of
equations (\ref{3ybey}) is consistent.
Note that the above considerations easily generalize to the case of
arbitrary $N$, which will be treated in the next subsection. In addition
to (\ref{braid}) and (\ref{idem}) we then will have to consider
(\ref{nnncom}) which leads to the condition
\begin{equation}
Y_{j, j+1} (s_1, s_2) Y_{k, k+1} (s_3, s_4) =
Y_{k, k+1} (s_3, s_4) Y_{j, j+1} (s_1, s_2) \qqd \mbox{for}\
|j - k| > 1 \qd,
\end{equation}
which is trivially satisfied.
Let us now impose periodic boundary conditions on the wave function
\bea
\psi(0,x_2,x_3;\sigma_1,\sigma_2,\sigma_3)&=&\psi(L,x_2,x_3;\sigma_1,\sigma_2,\sigma_3)\ ,\nonumber\\
\psi(1,x_2,x_3;\sigma_1,\sigma_2,\sigma_3)&=&\psi(L+1,x_2,x_3;\sigma_1,\sigma_2,\sigma_3)\ ,\nonumber\\
\psi(x_1,0,x_3;\sigma_1,\sigma_2,\sigma_3)&=&\psi(x_1,L,x_3;\sigma_1,\sigma_2,\sigma_3)\ ,\nonumber\\
\psi(x_1,1,x_3;\sigma_1,\sigma_2,\sigma_3)&=&\psi(x_1,L+1,x_3;\sigma_1,\sigma_2,\sigma_3)\ ,\nonumber\\
\psi(x_1,x_2,0;\sigma_1,\sigma_2,\sigma_3)&=&\psi(x_1,x_2,L;\sigma_1,\sigma_2,\sigma_3)\ ,\nonumber\\
\psi(x_1,x_2,1;\sigma_1,\sigma_2,\sigma_3)&=&\psi(x_1,x_2,L+1;\sigma_1,\sigma_2,\sigma_3)\ .
\label{3pbc}
\eea
Inserting \r{BAWF3} into \r{3pbc} yields
\begin{equation}
A_{\sigma_{Q_1}\sigma_{Q_2}\sigma_{Q_3}}(k_{P_1},k_{P_2},k_{P_3})=
\exp(i k_{P_1}L)
A_{\sigma_{Q_2}\sigma_{Q_3}\sigma_{Q_1}}(k_{P_2},k_{P_3},k_{P_1})\ ,
\label{3pbcA}
\end{equation}
where $Q$ and $P$ are arbitrary permutations of $\{1,2,3\}$.
In terms of the auxiliary spin model \r{3pbcA} is expressed as
\bea
|k_{P_1},k_{P_2},k_{P_3}\rangle&=&
\exp(i k_{P_1}L)
\Pi^{(1,2)}\Pi^{(2,3)}|k_{P_2},k_{P_3},k_{P_1}\rangle\nonumber\\
&=&\exp(i k_{P_1}L)
\Pi^{(1,2)}\Pi^{(2,3)}Y^{2,3}(\sin k_{P_1},\sin k_{P_3})
Y^{1,2}(\sin k_{P_1},\sin k_{P_2})|k_{P_1},k_{P_2},k_{P_3}\rangle\nonumber\\
&=&\exp(i k_{P_1}L)
\Pi^{(1,2)}X^{2,3}(\sin k_{P_1},\sin k_{P_3})\Pi^{(1,2)}
X^{1,2}(\sin k_{P_1},\sin k_{P_2})|k_{P_1},k_{P_2},k_{P_3}\rangle\nonumber\\
&=&\exp(i k_{P_1}L)
X^{1,3}(\sin k_{P_1},\sin k_{P_3})
X^{1,2}(\sin k_{P_1},\sin k_{P_2})|k_{P_1},k_{P_2},k_{P_3}\rangle\ ,
\eea
where we have used the identities \r{ppp}. Using \r{xtau} we now
obtain in complete analogy with the 2-electron case
\beq
|k_{P_1},k_{P_2},k_{P_3}\rangle =e^{i Lk_{P_1}}\
\tau(\sin k_{P_1}|\{\sin k_{P_j}; j=1, \ldots,3\})\
|k_{P_1},k_{P_2},k_{P_3} \rangle \qd.
\label{pbc33}
\eeq
We again can use the results for the diagonalization of the
inhomogeneous transfer matrix $\tau(\sin k_{P_1}|\{\sin k_{P_j}
;j=1, \ldots,3 \})$
derived in Appendix \ref{section:XXX}. We need to distinguish two
cases, depending on the spins of the two electrons
(recall that we consider only states for which the number of
down spins not larger than the number of up spins, all other states are
obtained by using the spin-reversal symmetry).
\begin{itemize}
\item{} \underline{Three electrons with spin up}
Here the appropriate eigenvector of $\tau(\Lambda|\{\sin
k_{P_j};j=1,\ldots ,3 \})$ is the ferromagnetic state, with eigenvalue $1$.
The periodic boundary conditions for the case of three electrons with
spin up thus take the form
\begin{equation}
e^{ik_j L}=1\ ,\qquad j=1,2,3.
\end{equation}
These are precisely the Lieb-Wu equations \r{bak},\r{bas}
for the case $N=3$, $M=0$.
Using \r{Astates} we see that all amplitudes are trivial
\begin{equation}
A_{\sigma_1\sigma_2\sigma_3}(k_{P_1},k_{P_2},k_{P_3})= 1\ .
\end{equation}
The corresponding wave-function \r{BAWF3} coincides with Woynarovich's
result \r{wswf}.
\item{} \underline{two electrons with spin up, one with spin down}
Now the appropriate eigenvector of $\tau$ is found in the sector with
one overturned spin. From \r{eigen} its eigenvalue is given by
\begin{equation}
1-iU/[2(\sin k_{P_1}-\lambda_1+iU/4)] = \frac{\sin k_{P_1}-\lambda_1-iU/4}
{\sin k_{P_1}-\lambda_1+iU/4}\ ,
\label{eval}
\end{equation}
where $\lambda_1$ fulfills
\bea
1 & = & \prod_{j=1}^{3} {\frac
{\lambda_1 - \sin k_j + iU/4}
{\lambda_1 - \sin k_j - iU/4}} \quad .
\label{lw31}
\eea
Inserting \
r{eval} into \r{pbc33} we obtain the following
quantization conditions due to periodic boundary conditions
\bea
\exp(i L k_{P_1}) & = & {\frac {\lambda_1 - \sin k_{P_1} - iU/4}
{\lambda_1 - \sin k_{P_1} + iU/4} } \quad , \qd {\rm for} \, P \in
S_3\ .
\label{lw32}
\eea
Equations \r{lw31} and \r{lw32} are precisely the Lieb-Wu equations
\r{bak},\r{bas} for the case $N=3$ and $M=1$.
An explicit expression for the amplitudes again is obtained from
\r{Astates} and \r{1spin}, with the result
\begin{equation}
A_{\sigma_1\sigma_2\sigma_3}(k_{P_1},k_{P_2},k_{P_3})=
\prod_{j=1}^{y-1} \left({\lambda_1 - \sin k_{P_j} - {iU\over
4}\over {\lambda_1 - \sin k_{P_j} + {iU\over 4}}}\right){1\over\lambda_1 -
\sin k_{P_y} + {iU\over 4}}\ ,
\label{aexpl3}
\end{equation}
where $y$ is the position of the down spin in the sequence
$\sigma_1\sigma_2\sigma_3$.
Inserting \r{aexpl3} into \r{BAWF3} we obtain \r{wswf} for the case
$N=3$, $M=1$.
\end{itemize}
\subsection{\boldmath $N$ electrons}
It is now clear how to generalize the above results to the case of $N$
electrons. The Bethe Ansatz for the solution $\psi$ of the
Schr\"odiger equation \r{SG} in the sector $Q$ with
$x_{Q_1}\leq x_{Q_2}\leq \ldots\leq x_{Q_N}$ is
\beq
\psi(x_1, \ldots,x_N; \sigma_1,\ldots,\sigma_N) =
\sum_{P \in S_N}\sign(PQ)
A_{\sigma_{Q_1}\ldots\sigma_{Q_N}}(k_{P_1},\ldots,k_{P_N})
\exp(i \sum_{j=1}^{N} k_{Pj}x_{Qj}) \quad .
\label{BAWFN}
\eeq
Substituting \r{BAWFN} into \r{SG} for the case $x_j\neq x_k, (j,
k=1,\ldots,N; j\neq k)$ we obtain the following expression for the energies
\beq
E = -2\sum_{j=1}^N \cos k_j -\frac{NU}{2}+\frac{UL}{4}\qd.
\eeq
Using the single valuedness of the wave function and solving the matching
conditions at the sector boundaries i.e. the Schr\"odinger equation
for the cases where two of the coordinates coincide, we obtain
the following set of equations
\bea
A_{\sigma_{Q'_1}\ldots\sigma_{Q'_N}}(k_{P'_1},\ldots,k_{P'_N})&=&
\sum_{\tau_1,\tau_2}
S^{\sigma_{Q_j}\tau_1}_{\sigma_{Q_{j+1}}\tau_2}(k_{P_j},k_{P_{j+1}})\
A_{\sigma_{Q_1}\ldots
\sigma_{Q_{j-1}}\tau_1\tau_2\sigma_{Q_{j+2}}\ldots\sigma_{Q_N}}
(k_{P_1},\ldots,k_{P_N})\ ,
\label{Nasa}
\eea
where $Q$ and $P$ are arbitrary permutations and
$Q'=Q(j,j+1)$,$P'=P(j,j+1)$.
In terms of the auxiliary spin model \r{Nasa} reads
\beq
|k_{P'_1},\ldots,k_{P'_N} \rangle =
Y^{j,j+1}(\sin k_{P_{j}}, \sin k_{P_{j+1}})
|k_{P_1},\ldots,k_{P_N} \rangle \quad .
\label{Nybey}
\eeq
The mutual consistency of equations \r{Nybey} follows from \r{yyy2}
and \r{yinv}.
We now impose periodic boundary conditions on the wave function
\bea
\psi(x_1,\ldots,x_{j-1},0,x_{j+1},\ldots,x_N;\sigma_1,\ldots,\sigma_N)&=&
\psi(x_1,\ldots,x_{j-1},L,x_{j+1},\ldots,x_N;\sigma_1,\ldots,\sigma_N)\
,\nonumber\\
\psi(x_1,\ldots,x_{j-1},1,x_{j+1},\ldots,x_N;\sigma_1,\ldots,\sigma_N)&=&
\psi(x_1,\ldots,x_{j-1},L+1,x_{j+1},\ldots,x_N;\sigma_1,\ldots,\sigma_N)\
,
\label{Npbc}
\eea
where $j=1,\ldots, N$. Inserting \r{BAWFN} into \r{Npbc} yields
\begin{equation}
A_{\sigma_{Q_1}\ldots\sigma_{Q_N}}(k_{P_1},\ldots,k_{P_N})=
\exp(i k_{P_1}L)
A_{\sigma_{Q_2}\ldots\sigma_{Q_N}\sigma_{Q_1}}(k_{P_2},\ldots,k_{P_N},k_{P_1})
\ ,
\label{NpbcA}
\end{equation}
where $Q,P\in S_N$ are arbitrary.
In terms of the auxiliary spin model \r{NpbcA} is expressed as
\bea
|k_{P_1},\ldots,k_{P_N}\rangle&=&
\exp(i k_{P_1}L)
\Pi^{(1,2)}\Pi^{(2,3)}\ldots\Pi^{(N-1,N)}
|k_{P_2},\ldots,k_{P_N},k_{P_1}\rangle\nonumber\\
&=&\exp(i k_{P_1}L)
\Pi^{(1,2)}\Pi^{(2,3)}\ldots\Pi^{(N-1,N)}
\left[\prod_{m=0}^{N-2} Y^{N-m-1,N-m}(\sin k_{P_1},\sin k_{P_{N-m}})\right]
|k_{P_1},\ldots,k_{P_N}\rangle\nonumber\\
&=&\exp(i k_{P_1}L)
X^{1,N}(\sin k_{P_1},\sin k_{P_{N}})
X^{1,N-1}(\sin k_{P_1},\sin k_{P_{N-1}})\ldots\nonumber\\
&&\times \ldots X^{1,3}(\sin k_{P_1},\sin k_{P_{3}})
X^{1,2}(\sin k_{P_1},\sin k_{P_{2}})
|k_{P_1},\ldots,k_{P_N}\rangle\ .
\eea
where we have used the identities \r{ppp}. Using \r{xtau} we now
obtain in complete analogy with the 2 and 3 electron cases
\beq
|k_{P_1},\ldots,k_{P_N}\rangle =e^{i Lk_{P_1}}\
\tau(\sin k_{P_1}|\{\sin k_{P_j}; j=1, \ldots,N \})\
|k_{P_1},\ldots,k_{P_N} \rangle \qd.
\label{pbc3N}
\eeq
Next we again use the results for the diagonalization of the
inhomogeneous transfer matrix $\tau(\sin k_{P_1}|\{\sin
k_{P_j};j=1,\ldots , N \})$ derived in Appendix \ref{section:XXX}.
We now need to distinguish $[N/2]+1$ cases ($[x]$ is the integer part
of $x$), corresponding to the possible values of $M$.
In the sector with $M$ down spins the eigenvalue of
$\tau(\sin k_{P_1}|\{\sin k_{P_j};j=1,\ldots , N \})$ is given by
\r{eigen}
\begin{equation}
\prod_{j=1}^M \frac{\sin k_{P_1}-\lambda_j-iU/4}
{\sin k_{P_1}-\lambda_j+iU/4}\ ,
\end{equation}
where the rapidities $\lambda_j$ fulfill
\bea
\prod_{l=1}^{N} {\frac
{\lambda_j - \sin k_l + iU/4}
{\lambda_j - \sin k_l - iU/4}}&=&
\prod_{\scriptstyle k=1\atop \scriptstyle k\neq j}^M
{\frac{\lambda_j - \lambda_k + iU/2}
{\lambda_j - \lambda_k - iU/2}} \quad .
\label{lwN1}
\eea
Inserting this result into \r{pbc3N} we finally obtain
\bea
\exp(i L k_{P_1}) & = & \prod_{j=1}^M{\frac {\lambda_j - \sin k_{P_1} - iU/4}
{\lambda_j - \sin k_{P_1} + iU/4} } \quad , \qd {\rm for} \, P \in
S_N\ .
\label{lwN2}
\eea
Equations \r{lwN1} and \r{lwN2} are precisely the Lieb-Wu equations
\r{bak},\r{bas} for the case of $N$ electrons, $M$ of which have spin down.
In order to obtain an explicit expression for the amplitudes we now
need to make use of the general result \r{stateswoy} for eigenstates
of the transfer matrix $\tau(\sin k_{P_1}|\{\sin k_{P_j};j=1,\ldots ,
N \})$ of the inhomogeneous Heisenberg model. We find
\begin{equation}
A_{\sigma_1\ldots\sigma_N}(k_{P_1},\ldots ,k_{P_N}) =
\sum_{\pi\in S_{M}} \ A_{\pi} \prod_{t=1}^M\left\lbrace
{1\over \lambda_{\pi_t}-\sin k_{P_{y_t}}+{iU\over 4}}
\prod_{s=1}^{y_t-1}\ {\lambda_{\pi_t}-\sin k_{P_s}-{iU\over 4}
\over \lambda_{\pi_t}-\sin k_{P_s}+{iU\over 4}}\right\rbrace\ ,
\end{equation}
where $1\le y_{_1}< y_{_2}< \ldots < y_{_M}\le N$ are the positions of
the down spins in the sequence $\sigma_1\ldots\sigma_N$ and $A_\pi$
is given by
\begin{equation}
A_\pi
= \prod_{1\le l< k\le M} \left({\lambda_{\pi_l} -
\lambda_{\pi_k}-{iU\over 2}\over
\lambda_{\pi_l} - \lambda_{\pi_k}}\right)\ \ .
\label{AAn}
\end{equation}
The resulting explicit expression for the wave function \r{BAWFN}
coincides with Woynarovich's result \r{wswf}.
\section{Inhomogeneous Heisenberg model}
\label{section:XXX}
\subsection{Algebraic Bethe Ansatz}
Our starting point is a lattice of $N$ spin-1/2's. The corresponding
Hilbert space is $V_1\otimes V_2\otimes\dots\otimes V_N$, where $V_j$
is isomorphic to $C^2$. We define the Pauli matrices ${\vec \tau}= (
\tau^{x}, \tau^{y},\tau^{z}) $ by
\beq
\tau^x=
\left(
\begin{array}{cc}
0 & 1 \nonumber \\
1 & 0 \nonumber \\
\end{array}
\right) \quad, \qd
\tau^y=
\left(
\begin{array}{cc}
0 & -i \nonumber \\
i & 0 \nonumber \\
\end{array}
\right) \quad, \qd
\tau^z=
\left(
\begin{array}{cc}
1 & 0 \nonumber \\
0 & -1 \nonumber \\
\end{array}
\right)
\eeq
and $\tau^\pm = \2 (\tau^x \pm i \tau^y)$.
The central object of the Quantum Inverse Scattering Method is the
$R$-matrix, which is a solution of the Yang-Baxter equation. For the
case of the spin-1/2 Heisenberg model it is of the form
\begin{equation}
{R(\lambda,\mu) =\left(\matrix{f(\mu,\lambda) &0&0&0\cr
0&g(\mu,\lambda)&1&0\cr 0&1&g(\mu,\lambda)&0\cr 0&0&0&f(\mu,\lambda)\cr
}\right) ,}
\label{R}
\end{equation}
where
\begin{equation}
{f(\mu,\lambda) = 1-{iU\over 2(\mu-\lambda) } \ ,\qquad
g(\mu,\lambda ) = -{iU\over 2(\mu - \lambda) } \ .}
\end{equation}
$R(\lambda,\mu)$ acts on the tensor product space $V_0\otimes V_0$, where
$V_0$ is isomorphic to $C^2$. The Yang-Baxter equation for $R$
is an equation on the space $V_0\otimes V_0\otimes V_0$ and can be
written as
\begin{equation}
R^{23}(\lambda,\mu)
R^{12}(\lambda,\nu)
R^{23}(\mu,\nu)
=R^{12}(\mu,\nu)
R^{23}(\lambda,\nu)
R^{12}(\lambda,\mu)\ ,
\label{rrr}
\end{equation}
where the superscript indicates in which spaces the $R$ matrix acts
nontrivially.
We now define an L-operator acting on the tensor product between a
``matrix-space'' $V_0$ and a ``quantum-space'' $V_n$, which is
identified with the Hilbert space over the $n^{\rm th}$ site of our
lattice of spins, by
\bea
L_n(\lambda) &=&
\frac{\lambda}{\lambda +iU/2} I +\frac{iU/2}{\lambda +iU/2}
\Pi^{(0,n)}\ ,\nonumber\\
&=&\frac{1}{\lambda+iU/2}
\left(
\begin{array}{cc}
\lambda + (1+\, \tau_n^z) iU/4 & \tau_n^{-}\ iU/2 \\
\tau_n^{+}\ iU/2 & \lambda + (1-\tau_n^z) iU/4
\end{array}
\right)\ .
\label{lop}
\eea
The Yang-Baxter equation \r{rrr} implies the following intertwining
relations for the $L$-operator
\begin{equation}
{R(\lambda,\mu) \left(L_n(\lambda)\otimes L_n(\mu)\right)
= \left(L_n(\mu)\otimes L_n(\lambda)\right) R(\lambda ,\mu) \ ,}
\label{intL}
\end{equation}
where the tensor product is between matrix spaces, i.e. \r{intL} is a
relation over the space $V_0\otimes V_0\otimes V_n$.
Next we note, that the intertwiner for the $L$-operator
(\ref{intL}) still holds, if we shift both spectral parameters $\lambda$
and $\mu$ by an arbitrary amount $\nu_n$, {\sl i.e.}
\begin{equation}
{R(\lambda ,\mu) \left(L_n(\lambda -\nu_n)\otimes
L_n(\mu -\nu_n)\right) = \left(L_n(\mu -\nu_n)\otimes L_n(\lambda
-\nu_n)\right) R(\lambda ,\mu) \ .}
\label{intLinh}
\end{equation}
We now construct an {\sl inhomogeneous monodromy matrix} in the
following way
\begin{equation}
{T(\mu|\{a_j\}) = L_N(\mu-a_N)L_{N-1}(\mu-a_{N-1})\ldots L_1(\mu-a_1) =
\left(\matrix{A(\mu)&B(\mu)\cr C(\mu)&D(\mu)\cr}\right) \ .}
\label{intT}
\end{equation}
Here $a_1,\ldots ,a_N$ are $N$ arbitrary complex constants.
The intertwiner \r{intLinh} can be lifted to the level of the monodromy
matrix
\begin{equation}
{R(\lambda ,\mu) \left(T(\lambda|\{a_j\})\otimes T(\mu|\{a_j\})\right)
= \left(T(\mu|\{a_j\})\otimes T(\lambda|\{a_j\})\right) R(\lambda ,\mu) \ .}
\label{rtt}
\end{equation}
By tracing (\ref{intT}) over the matrix space $V_0$ one then finds
that the {\sl transfer matrices}
\begin{equation}
\tau(\mu|\{a_j\})={\rm Tr}_0 (T(\mu|\{a_j\}))=A(\mu)+D(\mu)
\label{transferm}
\end{equation}
commute for any values of spectral parameter $\mu$, {\sl i.e.}
$[\tau(\mu|\{a_j\}),\tau(\nu|\{a_j\})]=0$. That implies that
the transfer matrix is the generating functional of an infinite number
of mutually commuting conserved quantum operators (via expansion in
powers of $\mu$).
Eigenstates of the transfer matrix are constructed by means of the
Algebraic Bethe Ansatz. Starting point is the choice of a {\sl
reference state}, which is a trivial eigenstate of $\tau(\mu|\{
a_j\})$.
We choose the saturated ferromagnetic state
\begin{equation}
|0\rangle = |\uparrow\up\uparrow\ldots\uparrow\rangle = \otimes_{n=1}^N |\uparrow\rangle_n\ .
\end{equation}
The action of the $L$-operator (\ref{lop}) on $|\uparrow\rangle_n$ can be
easily calculated and implies the following actions of the matrix
elements of the monodromy matrix
\bea
A(\mu)|0\rangle &=& a(\mu)|0\rangle\ ,\quad a(\mu) = 1\ ,\nonumber\\
D(\mu)|0\rangle &=& d(\mu)|0\rangle\ ,\quad d(\mu) =
\prod_{j=1}^N \frac{\mu-a_{j} }{\mu-a_{j} + iU/2 } \ ,\nonumber\\
C(\mu)|0\rangle &=&0\ ,\qquad B(\mu)|0\rangle \neq 0\ .
\label{AD}
\eea
{From} (\ref{AD}) we see that $B(\lambda)$ play the role of creation
operators. Acting with $B$ corresponds to flipping a spin.
States with $M$ down spins can be constructed as
\begin{equation}
F(\lambda_1,\ldots ,\lambda_M)=\prod_{j=1}^M (-2i/U)
B(\lambda_j -iU/4)|0\rangle \ ,
\label{states}
\end{equation}
where we have shifted the spectral parameters and introduced a
particular normalization for later convenience.
The requirement that the states (\ref{states}) ought to be eigenstates
of the transfer matrix puts constraints on the values $\lambda_n$: the
set $\{\lambda_j\}$ must be a solution of the following system of
Bethe Ansatz equations
\begin{equation}
\label{bae}
\prod_{k=1}^N \frac{\lambda_j-a_{k}+iU/4 }{\lambda_j-a_{k} - iU/4 }
=\prod_{\scriptstyle l=1\atop \scriptstyle l\neq j}^M
\frac{\lambda_j-\lambda_l+iU/2}{\lambda_j-\lambda_l-iU/2}\ ,
\qd j=1,\ldots ,M\ .
\end{equation}
The corresponding eigenvalues of the transfer matrix are
\begin{equation}
\tau(\mu|\{a_j\})\ F(\lambda_1,\ldots ,\lambda_M)=
\left[a(\mu)\prod_{j=1}^M f(\mu,\lambda_j-iU/4)
+d(\mu)\prod_{j=1}^M f(\lambda_j-iU/4,\mu)\right]
F(\lambda_1,\ldots ,\lambda_M)\ .
\label{eigen}
\end{equation}
For our present purposes we need to consider the transfer matrix
evaluated at the first inhomogeneity. We find
\bea
\tau(a_1|\{a_j\}) &=& {\rm Tr}_0\left[
L_N(a_1-a_N)L_{N-1}(a_1-a_{N-1})\ldots L_2(a_1-a_2)\Pi^{(0,1)}\right]\nonumber\\
&=&
{\rm Tr}_0\left[\Pi^{(0,1)}\Pi^{(0,1)}
L_N(a_1-a_N)\Pi^{(0,1)}\Pi^{(0,1)}L_{N-1}(a_1-a_{N-1})
\Pi^{(0,1)}\Pi^{(0,1)}\ldots \Pi^{(0,1)}L_2(a_1-a_2)\Pi^{(0,1)}\right]\nonumber\\
&=&
{\rm Tr}_0\left[\Pi^{(0,1)}
X^{1,N}(a_1,a_N)X^{1,N-1}(a_1,a_{N-1})\ldots X^{1,2}(a_1,a_2)\right]\nonumber\\
&=&
X^{1,N}(a_1,a_N)X^{1,N-1}(a_1,a_{N-1})\ldots X^{1,2}(a_1,a_2)\ ,
\label{xtau}
\eea
where $X^{j,k}$ is defined in \r{xjk}. Here we have first used
the explicit form of the L-operator \r{lop} and then the identities
\begin{equation}
\Pi^{(j,k)}\Pi^{(j,n)}\Pi^{(j,k)}=\Pi^{(k,n)}\ ,\qquad
\Pi^{(j,k)}\Pi^{(j,k)}=I\ .
\label{ppp}
\end{equation}
\subsection{Explicit expressions for the eigenstates}
In this subsection we derive the following expression for the
eigenstates \r{states}
\begin{equation}
F(\Lambda_1,\ldots ,\Lambda_M)=
\sum_{\{y_{_i}\}}
\sum_{\pi\in S_{M}} \ A_{\pi} \prod_{t=1}^M\left\lbrace
{1\over \Lambda_{\pi_t}-a_{P_{y_t}}+{iU\over 4}}
\prod_{s=1}^{y_t-1}\ {\Lambda_{\pi_t}-a_{P_s}-{iU\over 4}
\over \Lambda_{\pi_t}-a_{P_s}+{iU\over 4}}\right\rbrace
\prod_{j=1}^M\tau^-_{y_j} |0\rangle\ ,
\label{stateswoy}
\end{equation}
where the summation extends over $1\le y_{_1}< y_{_2}< y_{_3}< ... <
y_{_M}\le N$, $P$ is a permutation of $N$ elements, and the $A_\pi$
is given by
\begin{equation}
A_\pi
= \prod_{1\le l< k\le M} {\Lambda_{\pi_l} -
\Lambda_{\pi_k}-{iU\over 2}\over
\Lambda_{\pi_l} - \Lambda_{\pi_k}}\ \ .
\label{AA}
\end{equation}
Our discussion closely parallels \cite{EKS92a}. The basic tool
for proving \r{stateswoy} is the ``$n$-site generalized model''
\cite{IK84,IKR87} (see also \cite{KBIBo} p.151f and p.171).
\vskip .5cm
\underline{Generalised two-site model.}
\vskip .5cm
Let us first divide the product of $L$-operators in \r{intT} into two
parts
\bea
T_{_I}(\Lambda) &=& L_n(\Lambda - a_{n})\ldots
L_2(\Lambda -a_{2})\ L_1(\Lambda - a_{1})\ ,\nonumber\\
T_{_{II}}(\Lambda) &=& L_N(\Lambda - a_{N})\ldots
L_{n+2}(\Lambda -a_{n+2})\ L_{n+1}(\Lambda - a_{n+1})\ .
\label{intT2}
\eea
Clearly we have
\begin{equation}
T(\Lambda) = T_{_{II}}(\Lambda)\ \ T_{_{I}}(\Lambda)\ \ .
\end{equation}
Both $T_{_{II}}(\Lambda) $ and $T_{_{I}}(\Lambda)$ are $2\times 2 $
matrices
\begin{equation}
T_{_{I}}(\Lambda) =
\left(\matrix{A_{_I}(\Lambda)&B_{_I}(\Lambda)\cr
C_{_I}(\Lambda)&D_{_I}(\Lambda)\cr} \right)\ \ ,\ \
T_{_{II}}(\Lambda) =
\left(\matrix{A_{_{II}}(\Lambda)&B_{_{II}}(\Lambda)\cr
C_{_{II}}(\Lambda)&D_{_{II}}(\Lambda)\cr} \right)\ \ .
\end{equation}
By construction the matrix elements of $T_{_{I}}(\Lambda)$ commute with
the matrix elements of $T_{_{II}}(\Lambda)$. The commutation relations
of the matrix elements of the same $T$-operator are as in \r{rtt}
\begin{equation}
R\left(\Lambda_1,\Lambda_2\right) \bigl(T_{\alpha}(\Lambda_1)\otimes
T_{\alpha}(\Lambda_2)\bigr) = \bigl(T_{\alpha}(\Lambda_2)\otimes
T_{\alpha}(\Lambda_1)\bigr) R\left(\Lambda_1,\Lambda_2\right)\ ,\quad
\alpha=I,II\ .
\end{equation}
The matrix elements of $T$ can be expressed in terms of the matrix
elements of $T_{_I}$ and $T_{_{II}}$, {\sl e.g.,}
\begin{equation}
B(\Lambda) = A_{_{II}}(\Lambda)\ B_{_{I}}(\Lambda) +
B_{_{II}}(\Lambda) \ D_{_{I}}(\Lambda)\ \ .
\end{equation}
It is also possible to express the vectors $\prod_{j=1}^M B(\Lambda_j)
\ |0\rangle$ in terms of $B_{_I}$ and $B_{_{II}}$. In order to do this we
will use that
\begin{equation}
C_{\alpha}(\Lambda)|0\rangle = 0\ \ ,\ \ A_{\alpha}(\Lambda)|0\rangle =
a_{\alpha}(\Lambda)|0\rangle\ \ ,\ \ D_{\alpha}(\Lambda)|0\rangle =
d_{\alpha}(\Lambda)|0\rangle\ ,\quad \alpha=I,II.
\end{equation}
Here
\begin{equation}
a_{_{I}}(\Lambda) = 1\ ,\ \ d_{_{I}}(\Lambda) =
\prod_{j=1}^{n}\frac{\Lambda - a_j}{\Lambda - a_j + iU/2}\ ,
\end{equation}
and
\begin{equation}
a_{_{II}}(\Lambda) = 1\ ,\ \ d_{_{II}}(\Lambda) =
\prod_{j=n+1}^N \frac{\Lambda - a_j}{\Lambda - a_j + iU/2} \ \ .
\end{equation}
In \cite{IK84,IKR87} it was proved that
\bea
\prod_{j=1}^{M} B(\Lambda_j)\ |0\rangle =
\sum_{{\cal S}_I, \, {\cal S}_{II}}\
\prod_{\Lambda_k^I \in {\cal S}_I}\
\prod_{\Lambda_m^{II} \in {\cal S}_{II}}
a_{_{II}}(\Lambda_k^I)\
d_{_{I}}(\Lambda_m^{II})\ f(\Lambda_k^{I},\Lambda_m^{II})\
B_{_{II}}(\Lambda_m^{II})\ B_{_{I}}(\Lambda_k^I)|0\rangle \ .
\label{genmod}
\eea
On the right hand side of \r{genmod} we have summations with respect
to partitions of the set of all $\Lambda_j$'s into two subsets ${\cal
S}_I=\{\Lambda^{I}_k\}$ and ${\cal S}_{II}=\{\Lambda^{II}_m\}$.
Here $k$ labels different $\Lambda$ in the subset
${\cal S}_I$ and $m$ labels different $\Lambda$ in the subset
${\cal S}_{II}$.
The above model is called generalised two-site model because $T$ is
represented as a product of two factors. This is not sufficient for
our purposes however. Let us therefore now consider the so-called
\vskip .5cm
\underbar{Generalised $k$-site model.}
\vskip .5cm
Let us represent $T$ as a product of $k$ factors
\begin{equation}
T(\Lambda) = T_{k}(\Lambda)\cdot\cdot\cdot T_{2}(\Lambda)\
T_{1}(\Lambda)\ \ .
\label{445}
\end{equation}
Here each $T_{\alpha}(\Lambda)$ is a string of $L$-operators like in
\r{intT2}. The commutation relations of the matrix elements of each of
the $T_{\alpha}(\Lambda)$ is given by the same intertwiner as in \r{rtt}.
The matrix elements of $T_{\beta}(\Lambda)$ commute with the matrix
elements of $T_{\alpha}(\Lambda)$ if $\alpha\ne\beta$ and like for the
two-site model we have
\begin{equation}
T_{\alpha}(\Lambda) = \left(\matrix{A_\alpha(\Lambda)&
B_\alpha(\Lambda)\cr
C_\alpha(\Lambda)&D_\alpha(\Lambda)\cr}\right) \ \ ,
\end{equation}
\begin{equation}
C_\alpha(\Lambda) |0\rangle = 0\ \ \ ,\ \ \ A_\alpha(\Lambda) |0\rangle =
a_\alpha(\Lambda) |0\rangle
\ \ \ ,\ \ \ D_\alpha(\Lambda) |0\rangle = d_\alpha(\Lambda) |0\rangle \ \ .
\label{abcd}
\end{equation}
By explicitly multiplying the matrices in \r{445} we can express
$B(\Lambda)$ in terms of matrix elements of the $T_\alpha(\Lambda)$.
Iteration of \r{genmod} leads to the following expression for the
eigenfunctions of the transfer matrix $T(\Lambda)$
\bea
\prod_{j=1}^{M} B(\Lambda_j)\ |0\rangle &=&
\sum_{{\cal S}_1, \dots, \, {\cal S}_k}
\left(\prod_{\alpha =1}^k\
\prod_{\Lambda_{m_\alpha} \in {\cal S}_\alpha}
B_{_{\alpha}}(\Lambda_{m_\alpha}^\alpha)
|0\rangle\right) \ \times\nonumber\\
&& \qqd \times \left(\prod_{1\le\alpha < \beta\le k}\
\prod_{\Lambda_{m_\alpha} \in {\cal S}_\alpha}\
\prod_{\Lambda_{k_\beta} \in {\cal S}_\beta}
a_{_{\beta}}(\Lambda_{m_\alpha}^\alpha)\
d_{_{\alpha}}(\Lambda_{k_\beta}^\beta)\
f(\Lambda_{m_\alpha}^\alpha,\Lambda_{k_\beta}^{\beta}) \right) .
\label{gm2}
\eea
Here the summation is with respect to the partitions of the set of all
$\Lambda_j$'s into $k$ disjoint subsets ${\cal S}_\beta, \beta =
1,\dots, k$. The index $m_\alpha$ enumerates different $\Lambda$
in the subset ${\cal S}_\alpha$ and the index $k_\beta$ enumerates
different $\Lambda$ in the subset ${\cal S}_\beta$. Equation \r{gm2}
was first proved in \cite{IKR87}.
We now consider the special case of the generalized $N$-site model,
where $N$ is the length of the underlying lattice. This means that
each factor $T_\alpha(\Lambda)$ in our generalised $N$-site model is
identified with an individual $L$-operator in \r{intT}.
The eigenvalues in \r{abcd} are given by
\begin{equation}
a_{\alpha}(\Lambda) = 1\ ,\qquad
d_{\alpha}(\Lambda) = \frac{\Lambda - a_\alpha}
{\Lambda - a_\alpha + iU/2}\ .
\end{equation}
The $B$-operators are given by
\begin{equation}
B_\alpha(\Lambda) = \frac{iU/2}{\Lambda - a_\alpha + iU/2}
\, \tau_\alpha^- \ ,
\end{equation}
and have the important feature that $B_\alpha (\Lambda_1) B_\alpha
(\Lambda_2) = 0$. This implies that each set ${\cal S}_\alpha$ in
\r{gm2} consists of maximally one element. We are now in the position
to write down explicit expressions for the eigenstates \r{states}.
\vskip .5cm
a) \underline{One overturned spin.}
\vskip .5cm
Let us first consider the eigenfunctions in the sector with one
overturned spin. Application of \r{gm2} yields
\begin{equation}
F( \Lambda ) = (- 2i/U)
B(\Lambda -iU/4)|0\rangle = \sum_{y=1}^{N}\sigma_y^-|0\rangle
{1 \over \Lambda - a_y + {iU\over 4}}
\prod_{i=1}^{y-1}
{\Lambda - a_i - {iU\over 4}\over {\Lambda - a_i + {iU\over 4}}}\ .
\label{1spin}
\end{equation}
Here all sets ${\cal S}_\alpha$ in \r{gm2} except one are empty. This
one (one-element) set is ${\cal S}_{y}= \{\Lambda-iU/4\}$.
\vskip .5cm
a) \underline{Two overturned spins.}
\vskip .5cm
Now application of \r{gm2} leads to
\bea
F(\Lambda_1, \Lambda_2) &=& \sum_{1\le y_{_1}< y_{_2}\le
N}\sum_{\pi\in S_2} \sigma_{y_1}^-\
\sigma_{y_2}^-|0\rangle \frac{1}{\Lambda_{\pi_1}-a_{y_1}+iU/4}
\frac{1}{\Lambda_{\pi_2}-a_{y_2}+iU/4}\times\nonumber\\
&& \qqd \times\left( \prod_{j=1}^{y_1-1}
\frac{\Lambda_{\pi_1} - a_{j} - {iU\over 4}}{\Lambda_{\pi_1} -
a_{j} + {iU\over 4}}\right)\left(\prod_{l=1}^{y_2-1}
\frac{\Lambda_{\pi_2} - a_{l} - {iU\over 4}}
{\Lambda_{\pi_2} - a_{l} + {iU\over 4}}\right)
f(\Lambda_{\pi_1},\Lambda_{\pi_2})\ .
\label{bb}
\eea
Here $\pi$ is a permutation of two elements $1,2$ and
\begin{equation}
f(\Lambda_{\pi_1},\Lambda_{\pi_2}) =
{\Lambda_{\pi_1}-\Lambda_{\pi_2} - {iU\over 2}\over
\Lambda_{\pi_1}-\Lambda_{\pi_2}}\ .
\end{equation}
In this case only two subsets ${\cal S}_\alpha$ are nonempty. Each
of them consists of one element (${\cal S}_{y_1} =
\{\Lambda_{\pi_1}-iU/4\}$ and ${\cal S}_{y_2} =
\{\Lambda_{\pi_2}-iU/4\}$).
Equation \r{bb} is of the desired form \r{stateswoy} if we identify
\begin{equation}
A_\pi = {{\Lambda_{\pi_1} - \Lambda_{\pi_2}-{iU\over 2}\over
\Lambda_{\pi_1} - \Lambda_{\pi_2}}}\ \ ,
\end{equation}
which is in complete agreement with \r{AA}.
\vskip.5cm
c) \underline{M overturned spins.}
\vskip .5cm
The result for 2 overturned spins generalizes straightforwardly to $M$
overturned spins. The nonempty subsets ${\cal S}_\alpha$ in \r{gm2}
(each of which consists of exactly one element) are ${\cal S}_{y_1} =
\Lambda_{\pi_1}-iU/4$, ${\cal S}_{y_2} = \Lambda_{\pi_2}-iU/4$, \dots,
${\cal S}_{y_M} = \Lambda_{\pi_M}-iU/4$, where $\pi$ is some
permutation of $M$ elements. A straightforward calculation then yields
\r{stateswoy} and \r{AA}.
\section{\boldmath The spectrum of three electrons with one-down spin
for $L=6$}
We present a complete list of eigenstates of the Hubbard Hamiltonian
(\ref{hamiltonian}) in section V for the case $N=3$ and $M=1$ for a
6-site system ($L=6$) and $U=5$. The energy levels are listed in
increasing order.
The energy eigenvalues obtained by direct numerical
diagonalization of the Hamiltonian (the Householder-QR method) and by
the Bethe ansatz method coincide within an error of $\CO(10^{-15})$.
In the table only the energy eigenvalues obtained by Bethe ansatz
are listed. The last digit for each numerical value has a rounding
error. The symbols $S$ and $P$ denote the spin and momentum of the
eigenstate, respectively.
There are 90 eigenstates for $N=3$ and $M=1$, as we enumerated in
section V.C.3. The numerical results shown in the table confirm the
completeness of the Bethe ansatz.
\bigskip
\begin{tabular}{|c|c|c|r|r|}
\hline
No. & Energy & $S$ & $P/(\pi/3)$ & types \\ \hline
1 & $-4.19862084914891$ & 1/2 & 5 & real $I_j= 0.5,-0.5,-1.5,
J= 0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ k_j= 0.542662224387082,
- 0.315837723840216, - 1.27402205174346,
\La_\alpha= 0.587983554411128$} \\ \hline
2 & $-4.19862084914891$ & 1/2 & 1 & real $I_j= 1.5, 0.5,-0.5,
J= -0.5$ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$k_j = 1.27402205174346, 0.315837723840216,
- 0.542662224387082, \La_\alpha= - 0.587983554411128$} \\ \hline
3 & $-4.00000000000000$ & 3/2 & 0 & quartet $I_j= 1,0,-1$ \\ \hline
4 & $-3.35402752146807$ & 1/2 & 4 & real $I_j= 0.5,- 0.5,- 1.5,
J= - 0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ k_j = 0.231206350487730, - 0.682137186480516,
-1.64346426640041, \La_\alpha = - 1.27426628748753$} \\ \hline
5 & $- 3.35402752146807$ & 1/2 & 2 & real $I_j= 1.5, 0.5,-0.5,
J= 0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ k_j= 1.64346426640041, 0.682137186480516,
-0.231206350487730, \La_\alpha= 1.27426628748753$} \\ \hline
6 & $- 2.63449090541439$ & 1/2 & 0 & real $I_j= 1.5,-0.5,-1.5,
J= 0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ k_j= 1.52991315867874, -0.279871354037999,
-1.25004180464074, \La_\alpha= 0.845079113394529$} \\ \hline
7 & $-2.63449090541439$ & 1/2 & 0 & real $I_j= 1.5, 0.5,-1.5,
J= -0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ k_j= 1.25004180464074, 0.279871354037999,
-1.52991315867874, \La_\alpha= -0.845079113394529$} \\ \hline
8 & $-2.21221243379665$ & 1/2 & 4 & real $I_j= 0.5,-0.5,-2.5,
J= 0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ k_j= 0.554943548673639, -0.307417851034944,
-2.34192080003189, \La_\alpha= 0.644785906059544$} \\ \hline
9 & $-2.21221243379665$ & 1/2 & 2 & real $I_j= 2.5, 0.5,-0.5,
J= -0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ k_j= 2.34192080003189, 0.307417851034944,
-0.554943548673640, \La_\alpha= -0.644785906059544$} \\ \hline
10 & $-2.17072407234717$ & 1/2 & 5 & real $I_j= 1.5,-0.5,-1.5,
J= -0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ k_j= 1.22622874924835, -0.660511074050219,
-1.61291522639472, \La_\alpha= -1.15790499577381$} \\ \hline
11 & $-2.17072407234717$ & 1/2 & 1 & real $I_j= 1.5, 0.5,-1.5,
J= 0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ k_j= 1.61291522639472, 0.660511074050219,
-1.22622874924835, \La_\alpha= 1.15790499577381$} \\ \hline
12 & $-2.00000000000000$ & 3/2 & 1 & quartet $I_j= 2,0,-1$ \\ \hline
13 & $-2.00000000000000$ & 3/2 & 5 & quartet $I_j= 1,0,-2$ \\ \hline
14 & $-2.00000000000000$ & 3/2 & 3 & quartet $I_j= 2,1,0$ \\ \hline
15 & $-2.00000000000000$ & 3/2 & 3 & quartet $I_j= 0,-1,-2$ \\ \hline
\end{tabular}
\clearpage
\begin{tabular}{|c|c|c|r|r|}
\hline
No. & Energy & $S$ & $P/(\pi/3)$ & types \\ \hline
16 & $-1.68312858421806$ & 1/2 & 3 & real $I_j= 0.5,-0.5,-2.5,
J= -0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ k_j= 0.261468569600159, -0.628328469155396,
-2.77473275403456, \La_\alpha= -0.993984117066770$} \\ \hline
17 & $-1.68312858421806$ & 1/2 & 3 & real $I_j= 2.5, 0.5,-0.5,
J= 0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ k_j= 2.77473275403456, 0.628328469155396,
-0.261468569600159, \La_\alpha= 0.993984117066770$} \\ \hline
18 & $-1.00000000000000$ & 3/2 & 2 & quartet $I_j= 2,1,-1$ \\ \hline
19 & $-1.00000000000000$ & 3/2 & 4 & quartet $I_j= 1,-1,-2$ \\ \hline
20 & $-1.00000000000000$ & 3/2 & 2 & quartet $I_j= 3,0,-1$ \\ \hline
21 & $-1.00000000000000$ & 3/2 & 4 & quartet $I_j= 3,1,0$ \\ \hline
22 & $-0.867143471169408$ & 1/2 & 3 & real $I_j= 0.5,-1.5,-2.5,
J= 0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ k_j= 0.515010200239397, -1.28773972583561,
-2.36886312799358, \La_\alpha= 0.460329439565352$} \\ \hline
23 & $-0.867143471169409$ & 1/2 & 3 & real $I_j= 2.5, 1.5,-0.5,
J= -0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ k_j= 2.36886312799358, 1.28773972583561,
-0.515010200239397, \La_\alpha= -0.460329439565352$} \\ \hline
24 & $-0.763936521983257$ & 1/2 & 2 & real $I_j= -0.5,-1.5,-2.5,
J= 0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ k_j= -0.404815536059732, -1.33874801493405,
-2.44522665379261, \La_\alpha= 0.071452164223840$} \\ \hline
25 & $-0.763936521983257$ & 1/2 & 4 & real $I_j= 2.5, 1.5, 0.5,
J= -0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ k_j= 2.44522665379261, 1.33874801493405,
0.404815536059732, \La_\alpha= -0.07.1452164223840$} \\ \hline
26 & $-0.724935621196484$ & 1/2 & 5 & real $I_j= 1.5, 0.5,-2.5,
J= -0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ k_j= 1.27598401405461, 0.318718889893781,
-2.64190045514499, \La_\alpha= -0.568959086784317$} \\ \hline
27 & $-0.724935621196484$ & 1/2 & 1 & real $I_j= 2.5,-0.5,-1.5,
J= 0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ k_j= 2.64190045514499, -0.318718889893781,
-1.27598401405461, \La_\alpha= 0.568959086784317$} \\ \hline
28 & $-0.633335023683704$ & 1/2 & 5 & real $I_j= 1.5,-0.5,-2.5,
J= 0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ k_j= 1.54060965024643, -0.274638622892597,
-2.31316857855043, \La_\alpha= 0.886033905628400$} \\ \hline
29 & $-0.633335023683705$ & 1/2 & 1 & real $I_j= 2.5, 0.5,-1.5,
J= -0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$k_j= 2.31316857855043, 0.274638622892597,
-1.54060965024643, \La_\alpha= -0.886033905628400$} \\ \hline
30 & $-0.441363635441783$ & 1/2 & 4 & real $I_j= 1.5,-0.5,-2.5,
J= -0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$k_j= 1.24802157741378, -0.602680979199234,
-2.73973570060774, \La_\alpha= -0.869103625350605$} \\ \hline
\end{tabular}
\clearpage
\begin{tabular}{|c|c|c|r|r|}
\hline
No. & Energy & $S$ & $P/(\pi/3)$ & types \\ \hline
31 & $-0.441363635441782$ & 1/2 & 2 & real $I_j= 2.5, 0.5,-1.5,
J= 0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$k_j= 2.73973570060774, 0.602680979199234,
-1.24802157741378, \La_\alpha= 0.869103625350605$} \\ \hline
32 & $-0.164932552352930$ & 1/2 & 0 & real $I_j= 1.5, 0.5,-2.5,
J= 0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$k_j= 1.61980144017864, 0.665436431311280,
-2.28523787148992, \La_\alpha= 1.18390417778921$} \\ \hline
33 & $-0.164932552352929$ & 1/2 & 0 & real $I_j= 2.5,-0.5,-1.5,
J= -0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$k_j= 2.28523787148992, -0.665436431311280,
-1.61980144017864, \La_\alpha= -1.18390417778921$} \\ \hline
34 & 0.00000000000000 & 3/2 & 0 & quartet $I_j= 2,0,-2$ \\ \hline
35 & 0.00000000000000 & 3/2 & 1 & quartet $I_j= 3,1,-1$ \\ \hline
36 & 0.041639350031250 & 1/2 & 2 & real $I_j= 0.5,-1.5,-2.5,
J= -0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$k_j= 0.242333631055016, -1.61486765220468,
-2.81625618363673, \La_\alpha= -1.16526625596158$} \\ \hline
37 & 0.041639350031250 & 1/2 & 4 & real $I_j= 2.5, 1.5,-0.5,
J= 0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$k_j= 2.81625618363673, 1.61486765220468,
-0.242333631055016, \La_\alpha= 1.16526625596158$} \\ \hline
38 & 0.599066446144297 & 1/2 & 1 & real $I_j= -0.5,-1.5,-2.5,
J= -0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$k_j= -0.702579425111547, -1.67294123942593,
-2.86046709144551, \La_\alpha= -1.39028884293989$} \\ \hline
39 & 0.599066446144296 & 1/2 & 5 & real $I_j= 2.5, 1.5, 0.5,
J= 0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$k_j= 2.86046709144551, 1.67294123942593,
0.702579425111547, \La_\alpha= 1.39028884293989$} \\ \hline
40 & 0.625362070788712 & 1/2 & 4 & real $I_j= 1.5,-1.5,-2.5,
J= 0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$k_j= 1.49854259024225, -1.26094870462919,
-2.33198898800626, \La_\alpha= 0.722114430777064$} \\ \hline
41 & 0.625362070788713 & 1/2 & 2 & real $I_j= 2.5, 1.5,-1.5,
J= -0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$k_j= 2.33198898800626, 1.26094870462919,
-1.49854259024225, \La_\alpha= -0.722114430777064$} \\ \hline
42 & 1.00000000000000 & 3/2 & 1 & quartet $I_j= 3,1,0$ \\ \hline
43 & 1.00000000000000 & 3/2 & 5 & quartet $I_j= 3,0,-1$ \\ \hline
44 & 1.00000000000000 & 3/2 & 1 & quartet $I_j= 2,1,-1$ \\ \hline
45 & 1.00000000000000 & 3/2 & 5 & quartet $I_j= 1,-1,-2$ \\ \hline
\end{tabular}
\clearpage
\begin{tabular}{|c|c|c|r|r|}
\hline
No. & Energy & $S$ & $P/(\pi/3)$ & types \\ \hline
46 & 1.24466526322010 & 1/2 & 3 & real $I_j= 1.5,-1.5,-2.5,
J= -0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$k_j= 1.23359323035324, -1.58513579306981,
-2.79005009087323, \La_\alpha= -1.05370338517746$} \\ \hline
47 & 1.24466526322010 & 1/2 & 3 & real $I_j= 2.5, 1.5,-1.5,
J= 0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$k_j= 2.79005009087323, 1.58513579306981,
-1.23359323035324, \La_\alpha= 1.05370338517746$} \\ \hline
48 & 1.26651528335317 & 1/2 & 0 & real $I_j= 2.5,-0.5,-2.5,
J= 0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$k_j= 2.65775114820892, -0.311801389555702,
-2.34594975865322, \La_\alpha= 0.614983937128585$} \\ \hline
49 & 1.26651528335317 & 1/2 & 0 & real $I_j= 2.5, 0.5,-2.5,
J= -0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$k_j= 2.34594975865322, 0.311801389555702,
-2.65775114820892, \La_\alpha= -0.614983937128585$} \\ \hline
50 & 1.55774931070832 & 1/2 & 5 & real $I_j= 2.5,-0.5,-2.5,
J= -0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$k_j= 2.31151260511968, -0.609456221165668,
-2.74925393515061, \La_\alpha= -0.901701357480205$} \\ \hline
51 & 1.55774931070832 & 1/2 & 1 & real $I_j= 2.5, 0.5,-2.5,
J= 0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$k_j= 2.74925393515061, 0.609456221165668,
-2.31151260511968, \La_\alpha= 0.901701357480206$} \\ \hline
52 & 2.00000000000000 & 3/2 & 0 & quartet $I_j= 3,2,1$ \\ \hline
53 & 2.00000000000000 & 3/2 & 0 & quartet $I_j= 3,-1,-2$ \\ \hline
54 & 2.00000000000000 & 3/2 & 2 & quartet $I_j= 3,1,-2$ \\ \hline
55 & 2.00000000000000 & 3/2 & 4 & quartet $I_j= 3,2,-1$ \\ \hline
56 & 2.59087887855288 & 1/2 & 4 & real $I_j= 2.5,-1.5,-2.5,
J= 0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$k_j= 2.60714264000049, -1.28680359174393,
-2.36753659945316, \La_\alpha= 0.468661303303178$} \\ \hline
57 & 2.59087887855288 & 1/2 & 2 & real $I_j= 2.5, 1.5,-2.5,
J= -0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$k_j= 2.36753659945316, 1.28680359174393,
-2.60714264000049, \La_\alpha= -0.468661303303178$} \\ \hline
58 & 3.00000000000000 & 1/2 & 3 & $\eta$-pairing $\otimes I_j= 0$
\\ \hline
59 & 3.24859982950560 & 1/2 & 4 & real $I_j= 2.5,-1.5,-2.5,
J= -0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ k_j= 2.29420005680492, -1.59212356246732,
-2.79647159673079, \La_\alpha= -1.07985902893246$} \\ \hline
60 & 3.24859982950560 & 1/2 & 2 & real $I_j= 2.5,1.5,-2.5,
J= 0.5 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ k_j= 2.79647159673079, 1.59212356246732,
-2.29420005680492, \La_\alpha= 1.07985902893246$} \\ \hline
\end{tabular}
\clearpage
\begin{tabular}{|c|c|c|r|r|}
\hline
No. & Energy & $S$ & $P/(\pi/3)$ & types \\ \hline
61 & 3.53450070903252 & 1/2 & 4 & complex $m=10.0, \ell= 6.0,
J^{'}= 1.5, \sigma= -1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$q= 2.18262982829371, \xi= 1.51998108466021,
k_3= 6.10671585537857, \Lambda = 1.96074971503998$} \\ \cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = 0.138234567206297\times 10^{-3}$,
$\Im \delta = 0.236218958183487\times 10^{-3}$} \\ \hline
62 & 3.53450070903252 & 1/2 & 2 & complex $m= 8.0, \ell= 1.0,
J^{'}=-1.5, \sigma= 1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ q= 4.10055547888588, \xi= 1.51998108466021,
k_3= 0.176469451801019, \Lambda = -1.96074971503998$} \\ \cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = -0.138234567211404\times 10^{-3}$,
$\Im \delta = 0.236218958176826\times 10^{-3}$ } \\ \hline
63 & 4.00000000000000 & 3/2 & 3 & quartet $I_j= 3,2,-2$ \\ \hline
64 & 4.00000000000000 & 1/2 & 4 & $\eta$-pairing $\otimes I_j= 1$ \\
\hline
65 & 4.00000000000000 & 1/2 & 2 & $\eta$-pairing $\otimes I_j=-1$ \\
\hline
66 & 4.13147838656489 & 1/2 & 5 & complex $m= 5.0, \ell= 1.0,
J^{'}= 1.5, \sigma= -1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ q= 2.23719190707960, \xi= 1.45368482839940,
k_3= 0.761603941823783, \Lambda = 1.77325189299191$} \\ \cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = 0.307921735414718\times 10^{-3}$,
$\Im \delta = 0.266758568093550\times 10^{-3}$} \\ \hline
67 & 4.13147838656489 & 1/2 & 1 & complex $m=13.0, \ell= 6.0,
J^{'}=-1.5, \sigma= 1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ q= 4.04599340009998, \xi= 1.45368482839940,
k_3= 5.52158136535580, \Lambda = -1.77325189299191$} \\ \cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = -0.307921735405836\times 10^{-3}$,
$\Im \delta = 0.266758568103764\times 10^{-3}$ } \\ \hline
68 & 4.33177213111493 & 1/2 & 5 & complex $m=11.0, \ell= 6.0,
J^{'}= 0.5, \sigma= -1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ q= 2.77491534511596, \xi= 1.10073629130283,
k_3= 5.96934237293065, \Lambda = 0.601275696875128$} \\ \cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = -0.273302194108371\times 10^{-2}$,
$\Im \delta = -0.199420060157429\times 10^{-2}$ } \\ \hline
69 & 4.33177213111493 & 1/2 & 2 & complex $m= 7.0, \ell= 1.0,
J^{'}=-0.5, \sigma= 1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ q= 3.50826996206362, \xi= 1.10073629130283,
k_3= 0.313842934248937, \Lambda = 0.601275696875128$} \\ \cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = -0.119981837180916\times 10^{1}$,
$\Im \delta = -0.199420060157118\times 10^{-2}$ } \\ \hline
70 & 4.57467498439465 & 1/2 & 0 & complex $m= 6.0, \ell= 1.0,
J^{'}= 0.5, \sigma= 1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ q= 2.88935479499487, \xi= 1.07299907022393,
k_3= 0.504475717189853, \Lambda = 0.411558289383258$} \\ \cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = -0.399317226400675\times 10^{-2}$,
$\Im \delta = 0.222938603299694\times 10^{-3}$ } \\ \hline
\end{tabular}
\clearpage
\begin{tabular}{|c|c|c|r|r|}
\hline
No. & Energy & $S$ & $P/(\pi/3)$ & types \\ \hline
71 & 4.57467498439465 & 1/2 & 0 & complex $m=12.0, \ell= 6.0,
J^{'}=-0.5, \sigma= -1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ q= 3.39383051218472, \xi= 1.07299907022393,
k_3= 5.778709589989730, \Lambda = 0.411558289383258$} \\ \cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = -0.819123406502514\times 10^{0}$,
$\Im \delta = 0.222938603298584\times 10^{-3}$ } \\ \hline
72 & 4.71197559752276 & 1/2 & 3 & complex $m= 9.0, \ell= 5.0,
J^{'}= 1.5, \sigma= -1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$q= 2.16082673394224, \xi= 1.54894763250639,
k_3= 5.10312449288489, \Lambda = 2.04356190829635$} \\ \cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = 0.892857558656424\times 10^{-4}$,
$\Im \delta = 0.211992582481946\times 10^{-3}$ } \\ \hline
73 & 4.71197559752276 & 1/2 & 3 & complex $m= 9.0, \ell= 2.0,
J^{'}=-1.5, \sigma= 1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$q= 4.12235857323734, \xi= 1.54894763250639,
k_3= 1.18006081429470, \Lambda = -2.04356190829635$} \\ \cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = -0.892857558563165\times 10^{-4}$,
$\Im \delta = 0.211992582494824\times 10^{-3}$ } \\ \hline
74 & 5.58600548103459 & 1/2 & 4 & complex $m=10.0, \ell= 5.0,
J^{'}= 0.5, \sigma= -1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ q= 2.72669142318074, \xi= 1.11641164129461,
k_3= 5.01859266560450, \Lambda = 0.683372926842314$} \\ \cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = -0.186687577849687\times 10^{-2}$,
$\Im \delta = -0.244844007284084\times 10^{-2}$ } \\ \hline
75 & 5.58600548103459 & 1/2 & 2 & complex $m= 8.0, \ell= 2.0,
J^{'}=-0.5, \sigma= 1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ q= 3.55649388399885, \xi= 1.11641164129461,
k_3= 1.26459264157509, \Lambda = 0.683372926842313$} \\ \cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = -0.136487897790614\times 10^{1}$,
$\Im \delta = -0.244844007284262\times 10^{-2}$ } \\ \hline
76 & 5.95823319001949 & 1/2 & 0 & complex $m= 6.0, \ell= 2.0,
J^{'}= 1.5, \sigma= -1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$q= 2.27376322866332, \xi= 1.41368501399761,
k_3= 1.73565884985295, \Lambda = 1.66056016294642$} \\ \cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = 0.455885386176691\times 10^{-3}$,
$\Im \delta = 0.245749596857303\times 10^{-3}$ } \\ \hline
77 & 5.95823319001949 & 1/2 & 0 & complex $m=12.0, \ell= 5.0,
J^{'}=-1.5, \sigma= 1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ q= 4.00942207851627, \xi= 1.41368501399761,
k_3= 4.54752645732663, \Lambda = -1.660560162946420$} \\ \cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = -0.455885386181354\times 10^{-3}$,
$\Im \delta = 0.245749596852640\times 10^{-3}$ } \\ \hline
78 & 6.00000000000000 & 1/2 & 5 & $\eta$-pairing $\otimes I_j= 2$ \\
\hline
79 & 6.00000000000000 & 1/2 & 1 & $\eta$-pairing $\otimes I_j=-2$ \\
\hline
80 & 6.03334376215914 & 1/2 & 1 & complex $m= 7.0, \ell= 2.0,
J^{'}= 0.5, \sigma= 1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ q= 2.96456730023144, \xi= 1.06124089913747,
k_3= 1.40124825791330, \Lambda = 0.288686041886590$} \\ \cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = -0.375435197170576\times 10^{-2}$,
$\Im \delta = 0.208597324485038\times 10^{-2}$ } \\ \hline
\end{tabular}
\clearpage
\begin{tabular}{|c|c|c|r|r|}
\hline
No. & Energy & $S$ & $P/(\pi/3)$ & types \\ \hline
81 & 6.03334376215914 & 1/2 & 5 & complex $m=11.0, \ell= 5.0,
J^{'}=-0.5, \sigma= -1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$q= 3.31861800694814, \xi= 1.06124089913747,
k_3= 4.88193704926629, \Lambda = 0.288686041886588$} \\ \cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = -0.573617731801462$,
$\Im \delta = 0.208597324485194\times 10^{-2}$ } \\ \hline
82 & 6.70841301401077 & 1/2 & 2 & complex $m= 8.0, \ell= 4.0,
J^{'}= 1.5, \sigma= -1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$q= 2.16393969303817, \xi= 1.54471884698899,
k_3= 4.04970102349645, \Lambda = 2.03142493862260$} \\ \cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = 0.956259360065381\times 10^{-4}$,
$\Im \delta = 0.215691965483877\times 10^{-3}$ } \\ \hline
83 & 6.70841301401077 & 1/2 & 4 & complex $m=10.0, \ell= 3.0,
J^{'}=-1.5, \sigma= 1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$q= 4.11924561414142, \xi= 1.54471884698899,
k_3= 2.23348428368314, \Lambda = -2.031424938622590$} \\ \cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = -0.956259360207490\times 10^{-4}$,
$\Im \delta = 0.215691965477882\times 10^{-3}$ } \\ \hline
84 & 7.00000000000000 & 1/2 & 0 & $\eta$-pairing $\otimes I_j=3$ \\
\hline
85 & 7.48332665113181 & 1/2 & 1 & complex $m= 7.0, \ell= 3.0,
J^{'}= 1.5, \sigma= -1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$ = 2.20081321450656, \xi= 1.49694330556893,
k_3= 2.92875642936307, \Lambda = 1.89535079698079$} \\ \cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = 0.187321491183834\times 10^{-3}$,
$\Im \delta = 0.252337664304214\times 10^{-3}$ } \\ \hline
86 & 7.48332665113181 & 1/2 & 5 & complex $m=11.0, \ell= 4.0,
J^{'}=-1.5, \sigma= 1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$q= 4.08237209267303, \xi= 1.49694330556893,
k_3= 3.354428877816510, \Lambda = -1.89535079698079$} \\ \cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = -0.187321491188275\times 10^{-3}$,
$\Im \delta = 0.252337664298441\times 10^{-3}$ } \\ \hline
87 & 7.59363119464461 & 1/2 & 3 & complex $m= 9.0, \ell= 4.0,
J^{'}= 0.5, \sigma= -1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$q= 2.74080575567842, \xi= 1.11156342668968,
k_3= 3.94316644941255, \Lambda = 0.659158265098385$} \\ \cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = -0.212814421543628\times 10^{-2}$,
$\Im \delta = -0.234942161399343\times 10^{-2}$ } \\ \hline
88 & 7.59363119464461 & 1/2 & 3 & complex $m= 9.0, \ell= 3.0,
J^{'}=-0.5, \sigma= 1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$q= 3.54237955150117, \xi= 1.11156342668968,
k_3= 2.34001885776704, \Lambda = 0.659158265098384$} \\ \cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = -0.131618838598134\times 10^{1}$,
$\Im \delta = -0.234942161399543\times 10^{-2}$ } \\ \hline
89 & 8.02701965828632 & 1/2 & 2 & complex $m= 8.0, \ell= 3.0,
J^{'}= 0.5, \sigma= 1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$q= 2.89722498813356, \xi= 1.07154335884761,
k_3= 2.58313043330567, \Lambda = 0.398665945793610$} \\
\cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = -0.401340764685176\times 10^{-2}$,
$\Im \delta = 0.414789194807641\times 10^{-3}$ } \\ \hline
90 & 8.02701965828632 & 1/2 & 4 & complex $m=10.0, \ell= 4.0,
J^{'}=-0.5, \sigma= -1.0 $ \\ \cline{2-5}
& \multicolumn{4}{|c|}{$q= 3.38596031904603, \xi= 1.07154335884761,
k_3= 3.70005487387392, \Lambda = 0.398665945793609$} \\ \cline{2-5}
& \multicolumn{4}{|c|}{$\Re \delta = -0.793318483940372$,
$\Im \delta = 0.414789194806753\times 10^{-3}$ } \\ \hline
\end{tabular}
\section{Conclusions}
In this article we considered fundamental questions concerning
the exact solution of the one-dimensional Hubbard model. In appendices
A and B we presented a detailed derivation of the wave function. We
studied solutions of the Lieb-Wu equations. We concentrated our
attention on $k$-$\La$ strings. Unlike $\La$ strings they were never
carefully studied in the literature before. We arrived to the
conclusion that, for fixed coupling and large lattice size, $k$-$\La$
strings deviate from their ideal positions in a controllable way. In
the infinite chain limit $k$-$\La$ strings approach their ideal
positions. We also reviewed the thermodynamics of the model. There are
two different approaches to thermodynamics: one is based on strings,
whereas the other one is not. Both approaches show convincing agreement
for the calculation of the bulk thermodynamic properties of the Hubbard
model. Passing to the zero temperature limit we obtained the dispersion
curves of all elementary excitations at zero magnetic field, below
half-filling.
We would like to conclude with pointing out some interesting open
problems:
\begin{enumerate}
\item
Perhaps the most fundamental open problem is the calculation of the
norm of the wave function (\ref{sectorq})-(\ref{wwfsa}). Formulae for
norms of wave functions are known for Bethe ansatz solvable models with
only one level of Bethe ansatz equations, like the Bose gas with delta
interaction \cite{Gaudin83,Korepin82} or the XXX and XXZ spin-$\2$
chains \cite{Korepin82} (see also \cite{KBIBo}). For these models the
norms are expressed as determinants of the Jacobians of the Bethe ansatz
equations. The calculation of the norm of the Bethe ansatz wave
functions may be considered as a first step towards the calculation of
determinant representations of correlation functions \cite{KBIBo}.
\item
Another interesting open problem which we already mentioned above is
the calculation of the finite temperature correlation length within
the quantum transfer matrix approach. This requires to calculate the
second largest eigenvalue of the quantum transfer matrix. One has to
overcome certain technical subtleties coming from the fact that the
Hubbard model is a model of Fermions. A formalism capable of
calculating the correlation length of integrable fermionic systems has
recently been developed \cite{USW98,SSSU99,Sakai99}.
\item
At zero temperature some of the correlation functions of the Hubbard
model show a power law decay. Conformal field theory\footnote{In the
context of condensed matter physics conformal field theory is
equivalent to Luttinger liquid theory \cite{Haldane81}.} naturally
describes these powers (the set of conformal dimensions) \cite{FrKo90}.
Still, there are open problems within the conformal approach. For
example, since the expansion of the lattice operators in terms of
conformal fields is not known explicitly, the resulting expressions
contain unknown amplitudes. Some of these amplitudes may actually
vanish. Recently \cite{EsFr99} the vanishing of the amplitudes
corresponding to density correlations for the half-filled model was
shown by use of the SO(4) symmetry. For a more thorough understanding
of correlation functions an identification of the operators which
are of interest for the Hubbard model with the standard operators of
conformal field theory would be important. Recent work on a scaling
limit of the Hubbard model \cite{Melzer95,WoFo97,WoFo99} may prove to
be useful in this context.
\item
The predictions of the conformal approach to correlation functions
are limited to large distances, corresponding to very low energies.
Theoretically as well as from the point of view of recent experiments
on quasi one-dimensional structures in solids (e.g.\ \cite{KimEtal96,%
KimEtal97}) it would be highly appreciable to have a method for the
calculation of correlation functions at all energy scales. This problem
was recently tackled in \cite{EsKo99} within the form factor approach
\cite{KaWe78,BKW79,Smirnov92}, which was originally designed for
integrable 1+1 dimensional quantum field theories. In \cite{EsKo99} it
was argued that the form factor approach might as well apply to the
half-filled Hubbard model, and a formula for the two-spinon form
factor of the spin-operator $S_j^+$ was presented. It would be
interesting to extend the result to form factors of electronic creation
and annihilation operators, as such kind of extension could be directly
applied to the interpretation of the angle-resolved photo emission
spectroscopy data of \cite{KimEtal96,KimEtal97}.
\item
Despite the progress in the understanding of the mathematical structure
of the Hubbard model, which was achieved over the past few years and
which we briefly discussed in the introduction, we still feel
uncomfortable with the present stage of our knowledge. Shastry's
$R$-matrix \cite{Shastry86b,Shastry88b,OWA87}, which is the key for our
present understanding of the algebraic structure behind the Hubbard
model, is unusual as compared to $R$-matrices of other integrable
models. It does not possess the so-called difference property, i.e. it
is not a function of the difference of the spectral parameters alone.
The $S$-matrix at half-filling \cite{EsKo94a}, on the other hand,
possesses the difference property and can therefore be associated with a
Y(su(2))$\oplus$Y(su(2)) Yangian \cite{UgKo94}. The precise relation
between $R$-matrix and $S$-matrix is only understood in the rather
simple situation of an empty band \cite{MuGo97a,MuGo98a}. Because of
the lack of the difference property we can neither find a boost
operator for the Hubbard model by the reasoning of \cite{Tetelman82}
nor can we associate a spectral curve with it.
Another problem is the dimension of the elementary $L$-operator related
to Shastry's $R$-matrix, which is $4 \times 4$ (rather than $3 \times
3$ as one could guess naively from the fact that the Bethe ansatz for
the Hubbard model has two levels). For this reason there are too many
candidates for creation and annihilation operators in the algebraic
Bethe ansatz \cite{Shastry88b,RaMa97,MaRa98}. Again this redundancy
has only been partially understood in the empty band case
\cite{MuGo98a}. The known algebraic Bethe ansatz \cite{RaMa97,MaRa98}
is of involved structure and hopefully will be simplified in the future.
\end{enumerate}
\noindent
{\bf Acknowledgment.} We would like to thank C. M. Hung for drawing
figures 1-3 for us. We are indebted to H. Frahm, N. Kawakami,
B. M. McCoy, A. Schadschneider, J. Suzuki, A. M. Tsvelik and
M. Takahashi for helpful and stimulating discussions. This work was
supported by the EPSRC (F.H.L.E), by the Deutsche Forschungsgemeinschaft
under grant numbers Go 825/2-1 (F.G.) and Kl 645/3 (A.K.) and by the
National Science Foundation under grant number PHY-9605226 (V.E.K).
A.K. acknowledges support by the Sonderforschungsbereich 341,
K\"oln-Aachen-J\"ulich.
\section{Introduction}
The Hubbard model was introduced as a simple effective model for the
study of correlation effects of d-electrons in transition metals
\cite{Gutzwiller63,Hubbard63} (see also \cite{Kanamori63}). It is
believed to provide a qualitative description of the magnetic
properties of these materials and the Mott metal-insulator
transition \cite{Gebhard97}. Despite of its appealing conceptual
simplicity rigorous results for the Hubbard model are rare.
The dimension of the underlying lattice is a crucial parameter.
Two of the most important theorems valid for arbitrary lattice
dimension are due to Nagaoka \cite{Nagaoka66} and to Lieb \cite{Lieb89}.
Nagaoka's theorem states, that creating a single hole in a half-filled
connected lattice for infinitely repulsive interaction renders the ground
state ferromagnetic. Lieb's theorem is valid for arbitrary finite
repulsion. It states that on a bipartite lattice at half-filling the
ground state has spin $S = \2 ||B| - |A||$, where $|B|$ ($|A|$) is the
number of sites in the $B$ ($A$) sublattice. For reviews of rigorous
results about the Hubbard model in arbitrary dimensions see
\cite{Lieb95,Tasaki98}. Some simplifications occur in the limit of
infinite lattice dimension \cite{MeVo89,Vollhardt89,Mueller-Hartmann89}.
However, most exact results have been obtained for the one-dimensional
lattice. This is because a complete set of eigenfunctions of the
Hubbard Hamiltonian is only known for this case. The one-dimensional
Hubbard model was solved by Lieb and Wu \cite{LiWu68}. They used the
nested Bethe ansatz discovered in \cite{Yang67,Gaudin67},
There exists a vast literature on the Hubbard model. Some of the most
significant results have been collected in two reprint volumes. The
volume \cite{Montorsi92} gives a general overview for the years
1963-1990\footnote{See also \cite{Rasetti91}}. For a review of the
history of the exact solution in one dimension including a rather
exhaustive list of references until about 1992 we refer the reader to
\cite{KoEsBo}.
Most of the references listed in \cite{KoEsBo} are based on the seminal
1968 paper \cite{LiWu68} by Lieb and Wu. In this paper the problem of
diagonalizing the Hamiltonian was reduced to solving a set of coupled
nonlinear equations known as the Bethe ansatz or Lieb-Wu equations. Lieb
and Wu calculated the ground state energy of the system. They showed
that the model at half-filling is an insulator for arbitrary positive
value of the coupling $U$. In other words, they showed that the
half-filled model undergoes a Mott transition at critical coupling
$U = 0$.
In 1972 Takahashi proposed a classification of the solutions to the
Lieb-Wu equations \cite{Takahashi72}, which is commonly referred to as
`Takahashi's string hypothesis'. Analogous classifications are used
in all models solvable by the Bethe Ansatz method (see
e.g. \cite{Takahashi71,FaTa84} for the case of the Heisenberg model and
\cite{TsWi83} for the case of the Anderson model, which bears certain
similarities with the Hubbard model). The string hypothesis is the
basis of many subsequent publications. In the paper \cite{Takahashi72},
Takahashi used it to obtain a set of nonlinear integral equations that
determines the thermodynamics of the Hubbard model. By solving these
equations in some limiting cases, he was able to calculate the
low temperature specific heat in \cite{Takahashi74}.
At the beginning of the 80's Woynarovich resumed the study of the
excitation spectrum of the Hubbard model \cite{Woynarovich82a,%
Woynarovich82b,Woynarovich83a,Woynarovich83b} which was started ten
years earlier \cite{Ovchinnikov70,Coll74}. He gave a detailed analysis
of the charge excitations at half-filling \cite{Woynarovich82a} and was
the first to study gapped, spin-singlet charge excitations at
half-filling. These involve the first examples of excitations which in
the sequel will be called $k$-$\La$-string excitations
\cite{Woynarovich82b}. In his article \cite{Woynarovich82a}
Woynarovich presented the explicit form of the Bethe ansatz wave
function (see equations (\ref{sectorq})-(\ref{wwfsa}) below).
Since the publication of the reprint volume \cite{KoEsBo} there have
been several interesting developments. Two of the present authors
\cite{EsKo94a,EsKo94b} showed that the excitation spectrum at
half-filling in the absence of a magnetic field is given by the
scattering states of only four elementary excitations. Two of them carry
charge but no spin, the other two carry spin but no charge. These
elementary excitations are called holon, antiholon and spinon with spin
up or down, respectively. They form the fundamental representation of
SO(4). In the same articles \cite{EsKo94a,EsKo94b} the $S$-matrix of
the four quasiparticles was obtained. Thus there is now a complete and
satisfactory picture on the level of elementary excitations of
spin-charge separation in the one-dimensional Hubbard model at
half-filling. Spin and charge degrees of freedom can be excited
separately, but the corresponding quasiparticles do interact, albeit
weakly. The interaction is seen in the non-triviality of the
$S$-matrix. Spin-charge separation is one of the most interesting
properties of the one-dimensional Hubbard model. Recently experimental
evidence was found for the existence of spin-charge separation in quasi
one-dimensional materials \cite{KimEtal96,KimEtal97}.
Another interesting recent development is the calculation of the bulk
thermodynamic properties of the Hubbard model within the quantum
transfer matrix approach \cite{KlBa96,JKS98}. In contrast to the
traditional approach \cite{Takahashi72,Takahashi97,Takahashi99}, the
quantum transfer matrix approach leads to a finite number of non-linear
integral equations that determine the Gibbs free energy. This enables
high-precision numerical calculation of thermodynamic quantities such
as the charge and spin susceptibilities over the entire range of
doping, temperature and magnetic field. It further opens the
interesting perspective of calculating the correlation length at
arbitrary finite temperature.
It was shown in the papers \cite{CNC94a,CNC94b,PCCS97} how to use a
pseudo particle approach in order to obtain transport properties
(optical conductivity) of the one-dimensional Hubbard model.
There was also progress in the understanding of the algebraic structure
of the Hubbard model. Shiroishi and Wadati showed \cite{ShWa95}, that
the $R$-matrix, which was constructed earlier by Shastry
\cite{Shastry86b,Shastry88b,OWA87}, and which underlies the
integrability of the Hubbard model, satisfies the Yang-Baxter equation.
Martins and Ramos \cite{RaMa97,MaRa98} were able to construct a variant
of the algebraic Bethe ansatz for the Hubbard model. They obtained the
eigenvalue of the transfer matrix of the two-dimensional statistical
covering model (see also \cite{YuDe97}). This result was later used in
the quantum transfer matrix approach to the thermodynamics
\cite{JKS98}. Another interesting algebraic result was the discovery of
a quantum group symmetry of the Hubbard model on the infinite line.
The Hamiltonian is invariant under the direct sum of two Y(su(2))
Yangians \cite{UgKo94}. The relation of these Yangians to Shastry's
$R$-matrix was clarified in \cite{MuGo97a,MuGo98a}, where it was also
shown that the eigenstates of the Hubbard Hamiltonian on the infinite
interval, at zero density transform like irreducible representations of
one of the Yangians.
The purpose of this article is to give a pedagogical introduction
to the Bethe ansatz solution of the one-dimensional Hubbard model and
at the same time to fill some gaps in the previous literature. We
present a detailed account of the Bethe ansatz solution for periodic
boundary conditions and of the thermodynamics of the model. There are
two approaches to the thermodynamics. The approach of Takahashi
\cite{Takahashi72,Takahashi97,Takahashi99} relies on a string
hypothesis for the Hubbard model and is a natural generalization of
Yang and Yang's thermodynamic Bethe ansatz for the delta interacting
Bose gas \cite{YaYa69}. The second approach \cite{JKS98} is built on a
lattice path integral formulation of the partition function. We compare
both approaches and discuss their specific advantages. Special
attention is given to an aspect which, although fundamental, was
largely ignored in the previous literature, namely $k$-$\La$ strings.
$k$-$\La$ strings are spin-singlet bound states of electrons.
The Bethe ansatz for the one-dimensional Hubbard model \cite{LiWu68}
gives the eigenfunctions and eigenvalues of the Hubbard Hamiltonian
parametrized by two sets of quantum numbers $\{k_j\}$ and
$\{\la_l\}$, which are solutions of the Lieb-Wu equations (see
formulae (\ref{bak}), (\ref{bas}) below). The $k_j$ and $\la_l$ are
called charge momenta and spin rapidities, respectively. The Lieb-Wu
equations have finite and infinite solutions $k_j$, $\la_l$. They
should be considered separately, because they have different occupation
numbers. Every finite solution $k_j$ (or $\la_l$) can be occupied only
once. If two finite $k_j$ (or $\la_l$) coincide, the wave function
vanishes (see formulae (\ref{perm2}), (\ref{perm3}) and below). By
contrast, infinite $k_j$ (or $\la_l$) can be generally occupied more
than once. Later we shall explain that the multiplicities of occupation
of the infinite $k$'s and $\la$'s are given by the dimensions of the
representations of corresponding su(2) symmetry algebras. In order to
study this carefully the `regular' Bethe ansatz was defined in
\cite{EKS92a}. `Regular' means that all $k$'s and $\la$'s are finite.
They may be real or complex.
Takahashi's string hypothesis \cite{Takahashi72} is a statement about
the structure of the regular solutions of the Lieb-Wu equations in the
thermodynamic limit. Except for the real solutions (all $k_j$ and all
$\la_l$ real) there are solutions involving complex $k_j$ and $\la_l$.
The complex momenta and rapidities occur in two kinds of configurations,
which are symmetric with respect to the real axis. These configurations
are called strings. There are $\La$ strings involving only spin
rapidities and $k$-$\La$ strings, which involve charge momenta as
well. The $\La$ strings can be interpreted as bound states of magnons,
whereas the $k$-$\La$ strings describe spin singlet bound states of
electrons.
The $\La$ strings in the Hubbard model are similar to the $\La$ strings
in the isotropic Heisenberg spin chain, which have been extensively
studied in the literature \cite{Bethe31,Takahashi71,Gaudin71,EKS92c}.
Much less attention has been given to the $k$-$\La$ strings, which are
peculiar to the Hubbard model. They play an important role at half-%
filling, where the $k$-$\La$ two string enters the calculation of the
phase shift in the holon-holon scattering \cite{EsKo94a,EsKo94b}.
Holons are the lowest lying charge excitations of the Hubbard model.
At half-filling they have a gap. Below half-filling they are gapless,
however, whereas all $k$-$\Lambda$ strings lead to gapped excitations
in the thermodynamic limit (see section VI). Hence, $k$-$\La$ strings
below half-filling do not contribute to the low energy properties of
the model. What is their physical significance then? They do
contribute to the high temperature thermodynamic properties of the
Hubbard model. This is, in fact, the context in which they were first
introduced \cite{Takahashi72}. On the other hand, $k$-$\La$ strings
are interesting as a curious kind of excitations, which does not exist
in more simple integrable models.
According to the string hypothesis the $k$-$\La$ strings approach
certain ideal configurations as the number of lattice sites becomes
large. We call these configurations `ideal $k$-$\La$ strings'.
They are characterized by $m$ complex $\la$'s and $2m$ complex
$k$'s. The $\la$'s involved in an ideal $k$-$\La$ string have common
real part $\La'$, and their imaginary parts are $(m - 1) \frac{\i U}{4},
(m - 3) \frac{\i U}{4}, \dots, - (m - 1) \frac{\i U}{4}$. In the
repulsive case ($U > 0$) the $k$'s are given as
\beann
k^1 & = & \p - \arcsin(\La' + m \tst{\frac{\i U}{4}}) \qd, \\
k^2 & = & \arcsin(\La' + (m - 2) \tst{\frac{\i U}{4}}) \qd, \\
k^3 & = & \p - k^2 \qd, \\
& \vdots & \\
k^{2m - 2} & = & \arcsin(\La' - (m - 2) \tst{\frac{\i U}{4}})
\qd, \\
k^{2m - 1} & = & \p - k^{2m - 2} \qd, \\
k^{2m} & = & \p - \arcsin(\La' - m \tst{\frac{\i U}{4}}) \qd.
\eeann
This means that $\Re \sin (k^\a) = \La'$. Hence, the $\sin (k^\a)'s$
and the $\la$'s form a string in the complex plane.
In the thermodynamic limit, when the strings become ideal, the
variables describing their width can be eliminated from the Lieb-Wu
equations. We call the resulting set of equations discrete Takahashi
equations.
Some reformulation of the string hypothesis may be necessary before it
will be possible to achieve a rigorous mathematical proof. Yet, the
string hypothesis has passed many tests, and there is no doubt by now
that it describes the physics of the Hubbard model correctly. We
summarize our understanding of the issue and present several important
tests and consequences of the string hypothesis. We shall mostly
concentrate on the Hubbard model below half-filling, since the case of
half-filling was treated elsewhere \cite{EsKo94a,EsKo94b}.
Let us outline the plan of this article.
In section II we summarize known basic results about the Hubbard model.
This section contains a review of the Bethe ansatz solution of Lieb
and Wu \cite{LiWu68}. We present the wave function in the form given
by Woynarovich \cite{Woynarovich82a} and discuss its discrete
symmetries, namely the symmetries under permutations of electrons and
quantum numbers and the particle-hole and spin-reversal symmetries. This
leads to the notion of regular Bethe ansatz states. We proceed with
explaining the SO(4) symmetry \cite{HeLi71,Yang89,Pernici90,YaZh90,%
GoMu97b}, which is characteristic of the Hubbard model. Then we
introduce the discrete Takahashi equations. SO(4) symmetry and discrete
Takahashi equations are the prerequisites for the proof of completeness
of the Bethe ansatz \cite{EKS92b}, which is reviewed at the end of the
section.
Section III comprises a rigorous analytical study of the Lieb-Wu
equations for one down spin. For the sake of pedagogical clarity we
mostly focus on the cases of two and three electrons. This is our first
test of the string hypothesis. The result is positive: $k$-$\La$
strings do exist as solutions of the Lieb-Wu equations. They become
ideal in the thermodynamic limit. Furthermore, the counting of solutions
implied by the discrete Takahashi equations agrees in this case with
the counting obtained directly from the Lieb-Wu equations.
Section IV is devoted to the self-consistent solution of the Lieb-Wu
equations for three electrons and one down spin. Self consistency
arguments underly the derivation of the discrete Takahashi equations.
In the simple case considered in this section we can be more explicit.
We calculate the deviations of the rapidities and momenta (solutions
of the Lieb-Wu equations) from their ideal positions (solutions of the
discrete Takahashi equations). It turns out that these deviations
vanish exponentially in the thermodynamic limit. This is our second
positive test of the string hypothesis.
In section V we complement the analytical considerations of the
previous sections with numerical results. Numerical data based on the
Lieb-Wu equations are compared with data, which were obtained
independently of the Bethe ansatz by direct numerical diagonalization
of the Hamiltonian. We find perfect agreement between the two numerical
methods. The energy levels obtained by the two methods agree within a
numerical error of $\CO(10^{- 15})$. Our numerical study confirms the
completeness of the Bethe ansatz and the correctness of the counting of
the solutions implied by the string hypothesis. This is our third
successful test of the string hypothesis.
In section VI we review Takahashi's approach to the thermodynamics
of the Hubbard model \cite{Takahashi72}. We show that the dressed
energies of all $k$-$\La$ strings (bound states of electrons) follow
from Takahashi's integral equations (thermodynamic Bethe ansatz
equations) in the zero temperature limit. We can actually do better.
Starting from the thermodynamic Bethe ansatz equations and passing to
the zero temperature limit we obtain a complete classification of all
elementary excitations at zero magnetic field, below half-filling.
Takahashi derived his equations in order to calculate thermodynamic
quantities such as the specific heat or charge and spin
susceptibilities for the Hubbard model \cite{Takahashi72,%
Takahashi74,KUO89,UKO90}. Nowadays there is an independent method to
calculate these quantities, which does not rely on strings. It is
called the quantum transfer matrix method \cite{SuIn87,SAW90,JKS98}.
In section VII we compare the results of both methods and find that
they agree well. The string hypothesis also passes this significant
test. If the string hypothesis would miss one of the elementary
excitations, it should be visible in the thermodynamics.
Section VIII contains a brief conclusion and a list of interesting
open problems.
In appendix A we present a derivation of the Bethe ansatz wave function
for the Hubbard model. It can be seen from the derivation that charge
momenta and spin rapidities do not have to be real. To every solution
of the Lieb-Wu equations there corresponds a well defined periodic wave
function. This is particularly the case for the $k$-$\Lambda$ string
solutions.
In appendix B we derive the algebraic Bethe ansatz solution of the
inhomogeneous isotropic Heisenberg model, which is needed to construct
the Bethe ansatz wave function in appendix A.
Appendix C contains the tables of our numerical data for three
electrons and one down spin.
|
\section*{References}
|
\section{Introduction}
The energy scale of primordial inflation is supposed to be
hierarchically below the gravitational scale in order to produce
tiny fluctuations of the cosmic microwave background radiation
\cite{Lyt}.
The origin of the hierarchical scale of inflation
may be dimensional transmutation induced by gauge theory dynamics.%
\footnote{This is analogous to dynamical supersymmetry breaking,
which may provide the origin of the hierarchically small scale of
electroweak symmetry breaking.}
The gauge theory dynamics for inflation is expected to cause profound
effects on the evolution of the universe, since
the inflationary process may determine the vacuum of our universe
through dominant expansion of the spatial extension of the corresponding
inflationary vacuum.
In this paper,
we consider implications of this dynamical inflation
on vacuum selection of our universe,
in particular, on determination of spacetime symmetries
including its dimensions.
Now we expose a mixed model of
dilaton fixing
\cite{Iza}
and topological inflation
\cite{Kaw}%
\footnote{They are both $R$-invariant and so is the mixed model.}
as a simple example of dynamical inflation
which incorporates an inflationary selection of the vacuum
of our universe.
Let us consider a four-dimensional supersymmetric field theory
with a dilaton $\Phi$ and an inflaton $\phi$ supermultiplets
whose superpotential takes a form
\begin{eqnarray}
W = Xf(\Phi) + Zh(\Phi)(1 - \lambda \phi^2),
\label{SP}
\end{eqnarray}
where $X$ and $Z$ denote chiral superfields
and $f(\Phi)$ and $h(\Phi)$ are functions of
the dilaton superfield $\Phi$:
\begin{eqnarray}
f(\Phi) = f_1 e^{-a_1 \Phi} - f_2 e^{-a_2 \Phi}, \quad
h(\Phi) = h e^{-a \Phi}.
\end{eqnarray}
Here we set the gravitational scale equal to unity and regard it
as a universal cutoff in the theory.
We assume the couplings $f_1$, $f_2$, $h$ and $\lambda$ of order one and
$0 < a_1 \sim a_2 \ll a$.
This superpotential can be generated by dynamics of hypercolor
gauge interactions where the dilaton-dependent scales $e^{-a_i \Phi}$
and $e^{-a \Phi}$ with $a$'s related to $\beta$ functions of the gauge
interactions
arise from hyperquark condensations with the aid of a superpotential
considered in Ref.\cite{Iza}.
The potential in supergravity is given by
\begin{equation}
V = e^{K} (K_{AB} F^A F^{B*} - 3|W|^2),
\label{POT}
\end{equation}
where $K$ is a K{\" a}hler potential,
$K_{AB}$ denotes the inverse of the matrix
\begin{eqnarray}
{\partial^2 K \over \partial \phi_A \partial \phi_B^*},
\end{eqnarray}
with $\phi_A = X, Z, \Phi, \phi$,
and $F^A$ is given by
\begin{eqnarray}
F^A = {\partial W \over \partial \phi_A} + {\partial K \over \partial \phi_A}W.
\end{eqnarray}
For a generic K{\" a}hler potential, we have vacua of the model
given by
\begin{eqnarray}
F^A = W = 0,
\label{COND}
\end{eqnarray}
which realizes the unbroken supersymmetry and $R$ invariance.
Hence the vacuum expectation values of the dilaton are determined by
\begin{eqnarray}
f(\Phi) = 0,
\end{eqnarray}
which has runaway and fixed solutions,
\begin{eqnarray}
\Phi \rightarrow \infty
\end{eqnarray}
and
\begin{eqnarray}
\langle \Phi \rangle = {1 \over a_1-a_2}\ln {f_1 \over f_2},
\end{eqnarray}
respectively.%
\footnote{The gauge coupling constant $g$ is given by
$g^2 = 1/{\rm Re}\langle \Phi \rangle$.}
For the fixed dilaton, from Eq.(\ref{COND}), we find
$\langle X \rangle = \langle Z \rangle =0$
and
\begin{eqnarray}
\langle \phi \rangle = \pm \lambda^{-{1 \over 2}}.
\label{VAC}
\end{eqnarray}
Since $|e^{-a_i \langle \Phi \rangle}| \gg |e^{-a \langle \Phi \rangle}|$,
we may integrate out the superfields $X$ and $\Phi$ to obtain
an effective superpotential
\begin{eqnarray}
W_{eff} = Zh(\langle \Phi \rangle)(1 - \lambda \phi^2),
\end{eqnarray}
which results in topological inflation between the two vacua
Eq.(\ref{VAC}) for appropriate values of the couplings
\cite{Kaw}.
Namely, once the dilaton is fixed, the universe may evolve through
an inflationary stage.
On the other hand, the runaway vacuum
$\Phi \rightarrow \infty$
yields a free theory, which induces no inflation.
Under the chaotic initial condition
\cite{Lin}
of the dilaton field, the spatial extension of the vacuum
corresponding to the fixed dilaton dominates through the inflationary
process over that of the runaway vacuum
\cite{Iza}.
Although we have provided a specific model, for definiteness,
to demonstrate our point, the implications we consider
in the dynamical inflation scenario seem rather generic.
In the above model, the dilaton serves as an example of moduli%
\footnote{For an investigation of stringy modular cosmology, see
Ref.\cite{Ban}.}
which determine the form of low-energy field theory
describing our universe.
Moduli are not necessarily usual moduli fields but
may be any variables which parametrize the moduli space of
underlying high-energy theory.
The moduli space may be even disconnected, since the chaotic
initial condition of moduli allows highly excited states
which interpolate disconnected pieces of the moduli space.
For example, the spacetime dimension of the low-energy field theory
may be determined by moduli which describe the size of the internal
space of high-energy theory. Then the chaotic initial condition
of those moduli implies that the spacetime dimension is determined
through the inflationary process. If the corresponding inflation
is dynamical, the spacetime dimension is expected to be four or less,
since the dynamical scale due to gauge theory dynamics
may not be generated in five or more dimensions.
In four dimensions,
the presence of ${\cal N}=1$ supersymmetry (and not ${\cal N} > 1$)
may also originate from
inflationary selection of this type, since it seems suitable for
realizing scalar potentials which satisfy the slow-roll condition
for inflation
\cite{Lyt}.
\newpage
|
\section{Introduction}
Recent observations with Rossi X-ray Timing Explorer (RXTE) have revealed
kilohertz quasi-periodic oscillations (QPOs) in at least eighteen
low-mass X-ray binaries (LMXBs; see Van der Klis 1998a,b for a review;
also see Eric Ford's QPO web page at
http://www.astro.uva.nl/ecford/qpos.html for updated information).
These kHz QPOs are characterized by
their high levels of coherence (with $\nu/\Delta\nu$ up to $100$),
large rms amplitudes (up to $20\%$), and wide span of frequencies
($500-1200$ Hz). In almost all sources,
the X-ray power spectra show twin kHz peaks moving up and down in frequency
together as a function of photon count rate, with the separation frequency
roughly constant (The clear exceptions are Sco X-1 and 4U 1608-52,
van der Klis et al.~1997, Mendez et al.~1998a; see also
Psaltis et al.~1998. In Aql X-1, only a single QPO has been detected.).
Moreover, in several sources,
a third, nearly coherent QPO has been detected
during one or more X-ray bursts, at a frequency
approximately equal to the frequency difference between the
twin peaks or twice that value. (An exception is 4U 1636-53,
Mendez et al.~1998b.)
The observations suggest a generic beat-frequency model
where the QPO with the higher frequency is associated
with the orbital motion at some preferred orbital radius
around the neutron star, while the lower-frequency QPO results
from the beat between the Kepler frequency and the neutron star
spin frequency. It has been suggested that this preferred radius
is the magnetosphere radius (Strohmayer et al.~1996) or the sonic
radius of the disk accretion flow (Miller, Lamb and Psaltis 1998; see also
Klu\'zniak et al.~1990).
The recent observational findings (e.g., the variable frequency
separations for Sco X-1 and 4U 1608-52) indicate that the ``beat'' is
not perfect, so perhaps a boundary
layer with varying angular frequencies, rather than simply the neutron
star spin, is involved.
This paper is motivated by recent RXTE observation of the bright globular
cluster source 4U 1820-30 (Zhang et al.~1998), which has
revealed that, as a function of X-ray photon count rate, $\dot C$,
the twin QPO frequencies increase
roughly linearly for small photon count rates
($\dot C\approx
1600-2500$~cps)
and become independent of $\dot C$ for larger photon count
rates
($\dot C\approx 2500-3200$~cps).
(The QPOs become unobservable for still higher
count rates.) It was suggested that the $\dot C-{\rm independent}$
maximum frequency
($\nu_{\rm max}=1060\pm 20$~Hz) of the upper QPO corresponds to the orbital frequency
of the disk at the {\it inner-most stable circular orbit} (ISO)
as predicted by general relativity.
This would imply that the NS has mass of $2.2M_\odot$
(assuming a spin frequency of $275$~Hz). It has also been noted earlier
(Zhang et al.~1997), based on the narrow range of the maximal
QPO frequencies ($\nu_{\rm max}\approx 1100-1200$ Hz) in at least six sources
(which have very different X-ray luminosities), that
these maximum frequencies correspond to the Kepler
frequency at the ISO, which then implies that the
neutron star masses are near $2M_\odot$ (see also Kaaret et al.~1997).
The neutron star masses inferred from identifiying $\nu_{\rm max}$ with
the Kepler frequency at the ISO would, if confirmed,
be of great importance for constraining the properties of
neutron stars and for understanding the recycling processes
leading to the formation of millisecond pulsars.
However, while it is tempting to identify
$\nu_{\rm max}$ with the
the orbital frequency at the ISO, this
seemingly natural
interpretation may not be true.
One clue that this identification may not be correct is that
the inferred neutron star masses are substantially above the
masses of those neutron stars for which accurate determinations
are available (Thorsett \& Chakrabarty 1999) even though spin-up to
$\nu_s\sim 300\,{\rm Hz}$ only requires accretion of a very small
amount of material ($\llM_\odot$; \S 2).
The cause of the flattening of the $\nu_{\rm QPO}-\dot C$ correlation, and
the value of the maximum frequency, are still not understood
(and the existence of a plateau in $\nu_{\rm QPO}$ with
increasing $\dot M$ is debatable; e.g. Mendez et al.~1998c).
We suggest in \S 3 that the steepening of the magnetic
field, expected near the accreting neutron star, together with general
relativistic effect, naturally leads to the flattening in the $\nu_{\rm QPO}$
-$\dot M$ correlation. In \S 4 we advocate two alternative interpretations
of the maximum QPO frequency without invoking excessively large neutron star
masses.
\section{Possible Problems with Neutron Star Masses $\go 2~M_\odot$}
The most important concern for the inferred neutron star mass of
$\go 2M_\odot$ is an empirical one. LMXBs have long been thought
(e.g., Alpar et al.~1982) to be
the progenitors of binary millisecond radio pulsars. The recent
discovery of binary X-ray pulsar SAX J1808-3658 (with spin period 2.5~ms and
orbital period 2~hrs; Wijnands \& van der Klis 1998; Chakrabarty \&
Morgan 1998) appears to confirm this link.
Measurements of neutron star masses
in radio pulsar binaries give values in a narrow range
around $M\simeq 1.4M_\odot$; the data are consistent with a neutron star
mass function that is flat between $\go 1.1M_\odot$ and $\lo
1.6M_\odot$ at 95\% CL (Thorsett \& Chakrabarty 1999, Finn 1994).
The masses of neutron stars in X-ray binaries are also consistent
with $M\simeq 1.4\,M_\odot$ (e.g., van Kerkwijk et al.~1995).
Of particular interest is the $5.4$ ms recycled
pulsar B1855+09 with a white dwarf companion: this system
is thought to have gone through a LMXB phase (Phinney \& Kulkarni 1994),
and contains a neutron star with
$M=1.41\pm 0.10M_\odot$
(Thorsett \& Chakrabarty 1999; earlier Kaspi et al. 1994
estimated $M=1.50\pm^{0.26}_{0.14}M_\odot$).
The 23 ms pulsar PSR B1802-07, which is in a white dwarf
binary that is also thought to have gone through the LMXB
phase, has an inferred mass
$M=1.26\pm^{0.15}_{0.67}M_\odot$ ($95\%$ confidence;
Thorsett \& Chakrabarty 1999).
If $1.4~M_\odot$ is the mass of the neutron star immediately after
its formation in core collapse, then to make a $2.2\,M_\odot$
object would require accretion of material of at least $0.8\,M_\odot$.
Such large accretion mass may be problematic.
If we neglect torques on the star due to the interaction of its magnetic
field and the accretion disk, the added mass
needed to spin up the NS to a spin frequency $\nu_s=\Omega_s/(2\pi)$ is
\begin{equation}
\Delta M\simeq {I\Omega_s\over\sqrt{GMr_{\rm in}}}
\simeq 0.07M_\odot {I_{45}\over\sqrt{M_{1.4}r_6}}
\left({\nu_s\over 300\,{\rm Hz}}\right),
\label{eq:delm1}\end{equation}
where $I=10^{45}I_{45}$~g~cm$^2$ is the moment of inertia,
$M=1.4M_{1.4}M_\odot$ is the neutron star mass and
and $r_{\rm in}=10r_6$~km is the radius of the inner edge of the accretion
disk, which could correspond to either the stellar surface (radius $R$)
or the inner-most stable orbit (ISO) in the absence of a magnetic field
strong enough to influence the flow substantially
(see Cook et al.~1994).
When the neutron star magnetic field is strong enough, the inner radius
$r_{\rm in}$ corresponds to the Alfv\'en radius. (We note that the positions of
all known millisecond pulsars and binary pulsars in the $P-\dot P$
diagram for radiopulsars are consistent with spinup via accretion onto
neutron stars with dipolar surface fields $\go 10^{8-9}$ G.)
For magnetic accretion, we expect
\begin{equation}
I\dot\Omega_s=\dot M\sqrt{GMr_{\rm in}} f(\omega_s),
\end{equation}
where $\omega_s=\Omega_s/\Omega_K(r_{\rm in})$, with $\Omega_K(r_{\rm in})$
the Kepler frequency at $r_{\rm in}$. The dimensionless
function $f(\omega_s)$ includes
contributions to the angular momentum transport from magnetic stresses and
accreting material. It is equal to zero at some
equilibrium $\omega_s$, but the actual form of $f(\omega_s)$
depends on details of the magnetic field -- disk interaction.
Treating $r_{\rm in}$ as a constant, we find
\begin{eqnarray}
\Delta M &=&
{I\Omega_s\over\sqrt{GMr_{\rm in}}}\left[{1\over\omega_s}
\int_0^{\omega_s}\!\!{d\omega_s'\over f(\omega_s')}\right]\nonumber\\
&=&0.07M_\odot{I_{45}\over\sqrt{M_{1.4}r_6}}\biggl[\omega_s^{-1}
\ln\biggl({1\over 1-\omega_s}\biggr)\biggr]\biggl({\nu_s\over
300\,{\rm Hz}}\biggr)\nonumber\\
&=& 0.04M_\odot I_{45}M_{1.4}^{-2/3}\left[\omega_s^{-4/3}
\ln\left({1\over 1-\omega_s}\right)\right]\left({\nu_s\over 300~{\rm Hz}}\right)^{4/3},
\label{eq:delm2}
\end{eqnarray}
where, in the last two lines, we have adopted a simple functional form
$f(\omega_s)=1-\omega_s$; generically,
\begin{equation}
\Delta M\approx 0.07M_\odot{I_{45}\over\sqrt{M_{1.4}r_6}}
\biggl({\nu_s\over300\,{\rm Hz}}\biggr)\psi(\omega_s),
\end{equation}
where $\psi(\omega_s)\to 1$ for $\omega_s\ll\omega_{s,c}$, assuming that the torque
tends to zero at a critical value $\omega_s=\omega_{s,c}$.
Large $\Delta M$ is possible
if there is a lengthy phase of accretion with nearly zero net torque (e.g. accreting
$0.8M_\odot$ at a mean accretion rate of $\dot M=10^{17}{\rm g\,s^{-1}}$ would require about
400 Myr) following a much shorter phase of spin-up to $\omega_s\to\omega_{s,c}$
(e.g. accreting $0.05M_\odot$ at $\dot M=10^{17}{\rm g\,s^{-1}}$ would require 30 Myr).
If magnetic field decays during accretion (e.g. Taam \& van den Heuvel 1986,
Shibazaki et al. 1989), then the spin-up phase would have been even shorter.
(Spin diffusion due to alternating or stochastic epsiodes of spin-up and spin-down
[e.g. Bildsten et al. 1997, Nelson et al. 1997] might be allowed -- but constrained
-- in such a picture.) To accomodate masses as large as $2M_\odot$, these LMXBs must
be rather old and must have spun up rapidly at first, and then not at all for
$\go 90\%$ of their lifetimes. Gravitational radiation might provide a mechanism
for enforcing virtually zero net torque during the bulk of accretion (Bildsten
1998, Andersson et al. 1999). But equations (\ref{eq:delm1}) and (\ref{eq:delm2})
show that only very small $\Delta M$ is {\it required} to achieve $\nu_s\sim
300\,{\rm Hz}$, irrespective of the mechanism responsible for halting spin-up at
such frequencies.
\section{Steepening Magnetic Fields Near the Accreting Neutron Star}
We shall adopt, as a working hypothesis, that the upper QPO
frequency is approximately equal to the Kepler frequency at a
certain critical radius of the disk (Strohmayer et al.~1996;
Miller, Lamb \& Psaltis 1998; van der Klis 1998)\footnote{In the model
of Titarchuk et al.~(1998) , the QPO corresponds of vertical oscillation of
the disk boundary layer, but the oscillation frequency is equal to the local
Kepler frequency. Even in the ``non-beat'' frequency model of Stella and
Vietri (1998), the upper QPO frequency still corresponds to the orbital
frequency.} that is determined by
the combined effects of general relativity and stellar magnetic field.
For sufficiently strong magnetic fields, the disk may be
truncated near this radius, where matter flows out of the disk and is
funneled toward the neutron star. This critical radius then corresponds to the
usual Alfv\'en radius (Strohmayer et al.~1996). Even if the fields are relatively
weak ($10^7-10^8$~G) and the field geometry is such that
matter remains in the disk, the magnetic stress can still
slow down the orbital motion in the inner disk by taking away
angular momentum from the flow, and accreting gas then plunges toward
the star at supersonic speed -- a process that is also accelerated
by relativistic instability. In this case, the critical radius
would correspond to the sonic point of the flow (Lai 1998).
We neglect the possible role of radiative forces discussed by
Miller et al.~(1998). As emphasized by van der Klis (1998a),
the fact that similar QPO frequencies ($500-1200$~Hz)
are observed in sources with vastly different average luminosities
(from a few times $10^{-3}L_{\rm Edd}$ to near $L_{\rm Edd}$)
suggests that radiative effects cannot be the only factor
that induces the correlation of the QPO frequency and the X-ray flux
for an individual source.
Despite many decades of theoretical studies (e.g.,
Pringle \& Rees 1972; Lamb et al.~1973; Ghosh \& Lamb 1979;
Arons 1987; Spruit \& Taam 1990; Aly 1991; Sturrock 1991;
Shu et al.~1994;
Lovelace, Romanova \& Bisnovatyi-Kogan ~1995, ~1999; Miller \& Stone 1997), there remain considerable
uncertainties on the nature of the stellar magnetic field -- disk interactions.
Among the issues that are understood poorly are the
transport of magnetic field in the disk,
the configuration of the field threading the disk,
and the nature of outflows from the disk.
To sidestep these complicated questions,
we adopt a simple phenomenological prescription
for the vertical and azimuthal components of
the magnetic field on the disk,
\begin{equation}
B_z=B_0\left({R\over r}\right)^n,~~~~
B_\phi=-\beta B_z,
\label{eq:Bfield}
\end{equation}
where $B_\phi$ is evaluated at the upper surface of the
disk, and $\beta$ is the
azimuthal pitch angle of the field.
If we neglect the GR effect, the critical radius $r_{\rm in}$ is located
where the magnetic field stress dominates the angular momentum transport in the
disk, and it is approximately given by the condition
\begin{equation}
\dot M {d\sqrt{GMr}\over dr}=-r^2B_zB_\phi;
\label{eq:balance}\end{equation}
using the {\it ansatz} equation (\ref{eq:Bfield}), and assuming Keplerian
rotation (which may break down in a boundary layer near $r_{\rm in}$; e.g. Lovelace
et al. 1995), we find
\begin{equation}
r_{\rm in}=R\left(2\beta{B_0^2R^3\over\dot M\sqrt{GMR}}\right)^{2/(4n-5)}.
\end{equation}
and the Kepler frequency at $r_{\rm in}$ is
\begin{equation}
\nu_K(r_{\rm in})\propto \dot M^{3/(4n-5)}.
\label{eq:scale}\end{equation}
For a ``dipolar'' field configuration, $n=3$ and $\nu_K(r_{\rm in})\propto
\dot M^{3/7}$, as is well-known, but for smaller values of $n$, the
dependence steepens; for example, $\nu_K(r_{\rm in})\propto\dot M$ for a
``monopole'' field, $n=2$. The observed correlation $\nu_{\rm QPO}\propto\dot C$
may require $n<3$, although the relationship between $\dot C$ and $\dot M$
is unclear (Mendez et al. 1998c).
Unusual field topologies are possible as the disk approaches the surface
of the neutron star. Values of $n\neq 3$ (and even violation of power-law
scaling) might occur naturally, for open field configurations, which may be
prevalent because of differential rotation between the star and the disk
(e.g. Lovelace et al. 1995). MHD winds driven off a disk could
also result in $n\neq 3$ (e.g. Lovelace et al.
1995, Blandford \& Payne 1982). Disks that are fully
(Aly 1980, Riffert 1980) or partially (Arons 1993) diamagnetic will also
have non-dipolar variation in field strength near their inner edges
(see also \S 3.1 below). None of these possibilities {\it requires}
the field to be substantially non-dipolar {\it at the stellar surface},
although for disks that penetrate close to the star (at $r_{\rm in}-R\lo R$)
any non-dipolar field components, if strong enough, would be significant.
A particular field configuration that could explain the observed variation
of $\nu_{\rm QPO}$ with $\dot C$ might have $n<3$ at moderate values of $r_{\rm in}$, leading
to a strong correlation between $\nu_{\rm QPO}$ and $\dot M$ (and hence $\dot C$).
As $\dot M$ rises, the disk approaches the star, and the field topology
could become more complex, resulting in additional, non-power-law radial
steepening of the field strength. As is argued below,
this could happen even if the field is dipolar at the surface of the star,
particularly if the disk is diamagnetic. This steepening of the field results
in a flattening of the $\nu_{\rm QPO}-\dot M$ relation. Additional flattening
results from incipient general relativistic instability at the inner
edge of the disk.
\subsection{A Specific Ansatz: Diamagnetic Disk}
An illustration of the field steepening discussed above
is as follows. Consider a vacuum dipole field produced by the star
$|B_z|=\mu/r^3$ (in the equatorial plane perpendicular to the dipole axis).
Imagine inserting a diamagnetic disk
in the equatorial plane with inner radius $r_{\rm in}$. Flux conservation
requires $\pi (r_{\rm in}^2-R^2)|{\bar B_z}|=2\pi \mu/R$, which gives
the mean vertical field inside between $R$ and $r_{\rm in}$:
\begin{equation}
|\bar B_z(r_{\rm in})|={2\mu\over R(r_{\rm in}^2-R^2)}.
\label{eq:Baver}\end{equation}
This field has scaling $|\bar B_z|\propto 1/r^2$ for large $r$, which would
result in $\nu_K(r_{\rm in})\propto \dot M$ (see
eq.~[\ref{eq:scale}]), and stiffens as the disk approaches the stellar
surface.
The actual field at $r=r_{\rm in}$ is difficult to calculate. Aly (1980)
found the magnetic field of a point dipole in the presence of
a thin diamagnetic disk (thickness $H\ll r$ at radius $r$),
and demonstrated that the field strength
at $r_{\rm in}$ is enhanced by a factor
$\sim (r_{\rm in}/H)^{1/2}$.
(See also Riffert 1980 and Arons 1993.)
However, the situation is different for a
finite-sized dipole (a conducting sphere of radius $R$) in the presence of
a diamagnetic disk. This can be seen by
considering a simpler problem, where we replace the disk
by a diamagnetic sphere (with radius $r_{\rm in}$)\footnote{ In replacing the
disk with a spherical surface, we lose the square-root divergence found by
Aly (1980) for infinitesmal $H/r$. But note that for small fields, the
disk penetrates near the star, and $H$ may not be {\it very} small compared
with $r_{\rm in}-R$. In assuming a point dipole, Aly (1980) (and Riffert 1980)
exacerbated the divergence, and their results probably apply only when
$r_{\rm in}\gg R$.}.The magnetic field at
radius $r$ (between $R$ and $r_{\rm in}$) is given by (in spherical
coordinates with the magnetic dipole along the $z$-axis):
\begin{eqnarray}
B_r(r,\theta) &=&
\left(-{2\mu\over r_{\rm in}^3}+{2\mu\over r^3}\right){\cos\theta\over
1-\alpha^3},\\
B_\theta (r,\theta) &=&
\left({2\mu\over r_{\rm in}^3}+{\mu\over r^3}\right){\sin\theta\over
1-\alpha^3},
\end{eqnarray}
where $\alpha=R/r_{\rm in}$. Thus the vertical magnetic field at the inner edge
of the disk ($r=r_{\rm in}$) is
\begin{equation}
|B_z(r_{\rm in})|={3\mu\over r_{\rm in}^3-R^3}.
\label{eq:moddipole}\end{equation}
We see that the magnetic field steepens as $r_{\rm in}$ approaches the stellar
surface. In reality, some magnetic field will penetrate the disk because of
turbulence in the disk and Rayleigh-Taylor instabilities (Kaisig, Tajima
\& Lovelace 1992); however, some steepening of the field may remain.
Adopting the magnetic field {\it ansatz} (\ref{eq:Baver}) and
$B_\phi=-\beta B_z$, we can use (\ref{eq:balance})
to calculate $r_{\rm in}$; this gives
\begin{equation}
2\,b^2{x_c^{2.5}\over (x_c^2-1)^2}=1,
\label{eq:rinone}
\end{equation}
where $x_c=r_{\rm in}/R$ and
\begin{equation}
b^2={\beta B_0^2R^3\over\dot M\sqrt{GMR}}=
0.07\left(M_{1.4}^{-1/2}R_{10}^{5/2}\right)
\left({\beta B_7^2\over \dot M_{17}}\right),
\end{equation}
$\mu=B_0^2R^3/2$ ($B_0=10^7B_7$~G is the polar
field strength at the neutron star surface),
$M_{1.4}=M/(1.4\,M_\odot)$, $R_{10}=R/(10\,{\rm km})$,
and $\dot M_{17}=\dot M/(10^{17}\,{\rm g\,s}^{-1})$.
Alternatively, if we adopt (\ref{eq:moddipole}), we find
\begin{equation}
{9\over 2}\,b^2{x_c^{2.5}\over (x_c^3-1)^2}=1.
\label{eq:rintwo}
\end{equation}
Figure 1 shows the Kepler frequency at $r_{\rm in}$ as a function the
scaled mass accretion rate, $\dot M_{17}M_{1.4}^{1/2}R_{10}^{-5/2}
M_{17}/\beta B_7^2=0.07/b^2$. Clearly, for small $\dot M$, $\nu_K(r_{\rm
in})$ depends on $\dot M$ through a power-law, but the dependence weakens as
$\dot M$ becomes large, in qualitative agreement with the
observed $\nu_{\rm QPO}$-$\dot M$ correlation.
General relativistic effects also flatten
the $\nu_{\rm QPO}$-$\dot M$ relation, as we discuss next.
\subsection{General Relativistic Effects}
General relativity (GR) introduces two effects on the location of the
inner edge of the disk. First, the space-time curvature modifies
the vacuum dipole field. For example, in Schwarzschild metric, the locally
measured magnetic field in the equatorial plane is given by
\begin{equation}
B^{\hat\theta}={\mu\over r^3}\left[6y^3(1-y^{-1})^{1/2}\ln(1-y^{-1})
+{6y^2(1-y^{-1}/2)\over (1-y^{-1})^{1/2}}\right],
\end{equation}
where $y=rc^2/(2GM)$ (Petterson 1974; Wasserman \& Shapiro 1983).
The GR effect steepens the field only at small $r$. For $r=6GM/c^2-10GM/c^2$,
we find the approximate scaling $B^{\hat\theta}\propto r^{-3-\epsilon}$,
with $\epsilon \simeq 0.3-0.4$. We shall neglect such a small
correction to the dipole field given the much larger uncertainties
associated with the magnetic field -- disk interaction.
A more important effect of GR is that it modifies the the dynamics of
the accreting gas around the neutron star.
Without magnetic field, the inner edge of the disk
is given by the condition $d l_K/dr=0$, where $l_K$ is the
specific angular momentum of a test mass:
\begin{equation}
l_K=\left({GMr^2\over r-3GM/c^2}\right)^{1/2}.
\end{equation}
This would give the usual the ISO at $r_{\rm iso}=6GM/c^2$,
where no viscosity is necessary to induce accretion\footnote{When viscosity
and radial pressure force is taken into account, the flow is transonic,
with the sonic point located close to $r_{\rm iso}$.}.
Since magnetic fields take angular momentum
out of the disk, we can determine the inner edge of the disk using an
analogous expression\footnote{Note that in the limit of perfect conductivity,
it is possible to express the Maxwell stress tensor in terms of a
magnetic field four-vector
$\bf B$ that is orthogonal to the fluid velocity four-vector
$\bf U$ (e.g. Novikov \& Thorne 1973, pp. 366-367). The field
components $B_\phi$ and $B_z$
in eq. (\ref{eq:balance2}) and below are actually the projections
of $\bf B$ onto a local orthonormal basis (i.e. $B_\phi\to\vec{\bf e}_{\hat\phi}
\cdot{\bf B}$ and $B_z\to\vec{\bf e}_{\hat z}\cdot{\bf B}$)
even though we have retained the nonrelativistic notation for these
field components. No additional relativistic corrections are required with
these identifications understood.}
\begin{equation}
\dot M {dl_K\over dr}=-r^2B_zB_\phi,
\label{eq:balance2}\end{equation}
(see eq.~[\ref{eq:balance}]). In Lai (1998) it was shown that
this equation determines the limiting value of the sonic point
of the accretion flow (although a Newtonian pseudo-potential was
used in that paper). Adopting the magnetic field {\it ansatz} (\ref{eq:Baver}),
we find
\begin{equation}
2\,b^2{x_c^{2.5}\over (x_c^2-1)^2}=\left(1-{6GM\over c^2r_{\rm in}}\right)
\left(1-{3GM\over c^2r_{\rm in}}\right)^{-3/2}.
\label{eq:gr1}\end{equation}
Similarly, using (\ref{eq:moddipole}), we have
\begin{equation}
{9\over 2}\,b^2{x_c^{2.5}\over (x_c^3-1)^2}=\left(1-{6GM\over c^2r_{\rm
in}}\right)\left(1-{3GM\over c^2r_{\rm in}}\right)^{-3/2}.
\label{eq:gr2}\end{equation}
It is clear that for $b\gg 1$, eq.~(\ref{eq:gr1}) or (\ref{eq:gr2})
reduces to the Newtonian limit (see \S 3.1), while for $b=0$ we recover
the expected $r_{\rm in}=r_{\rm iso}=6GM/c^2$. For small $b$, the GR
effect can modify the inner disk radius signficantly.
In Fig.~1 we show the orbital frequency at $r_{\rm in}$
(as a function of the ``effective'' accretion rate) as obtained
from (\ref{eq:gr1}) and (\ref{eq:gr2}). We see that the GR effect induces additional
flattening in the correlation between $\nu_K(r_{\rm in})$ and $\dot M$
as $r_{\rm in}$ approaches $r_{\rm iso}$.
We emphasize the phenomenological nature of
eqs.~(\ref{eq:balance2})-(\ref{eq:gr2}): they are not derived from a
self-consistent MHD calculation, and take account of the dynamics
of the disk under a prescribed magnetic field configuration.
However, we believe that they indicate the combined effects
of dynamically altered magnetic field
and GR on the inner region of the accretion disk.
By measuring the correlation between the QPO frequency and the mass
accretion rate, one might be able to
constrain the magnetic field structure in accreting
neutron stars, and reach quantitative conclusions about
the nature of the interaction of the accretion disk and
magnetic field.
\section{Where are the QPOs Produced?}
Implicit in the discussion of magnetic fields and $\nu_{\rm QPO}$ in the
preceding sections were the assumptions that the QPO arises at
a radius outside the star that coincides with the inner radius
of the accretion disk. Here, we examine two ways
in which these assumptions might be violated, and show how the
relatively small measured values of $\nu_{\rm max}$ might be consistent
with neutron star masses near $1.4M_\odot$.
\subsection{Disk Termination at the Neutron Star Surface}
For the model discussed in \S 3, the steepening magnetic field
and general relativity produce the flattening in the correction between
the QPO frequency $\nu_{\rm QPO}=\nu_K(r_{\rm in})$ and the mass
accretion rate $\dot M$. But $\nu_{\rm QPO}$ becomes
truly independent of $\dot M$ only when $r_{\rm in}$ approaches
$r_{\rm iso}$ or the stellar radius $R$. It has been suggested (see \S 1)
that the $\dot M$-independent QPO frequency corresponds to
the Kepler frequency at $r_{\rm iso}$. But it is also possible
that the inner disk radius reaches the stellar surface, which is outside
the ISO,
as $\dot M$ increases. We note that observationally it is difficult
to distinguish the flattening of $\nu_{\rm QPO}$ and a true plateau.
It is not clear that the flattening feature at $\nu_{\rm QPO}\sim 1100$~Hz
observed in 4U 1820-30 (Zhang et al.~1998) corresponds the maximum
QPO frequency, but we shall assume it does and explore the consequences.
The maximum QPO frequency, $\nu_{\rm max}$,
is given by the orbital frequency at the larger of $r_{\rm iso}$ and $R$.
To linear order in $\nu_s$ (the spin frequency), the ISO is located at
$r_{\rm iso}=(6GM/c^2)(1-0.544\,a)$, and the orbital frequency at ISO is
\begin{equation}
\nu_K(r_{\rm iso})={1571\over M_{1.4}}(1+0.748\,a)~{\rm Hz},
\end{equation}
with the dimensionless spin parameter
\begin{equation}
a\simeq 0.099\, {R_{10}^2\over M_{1.4}}\left({\nu_s\over 300\,{\rm Hz}}
\right),
\end{equation}
where we have adopted $I=(2/5)\kappa MR^2$ for the moment of inertia of
the neutron star, with $\kappa\simeq 0.815$ (appropriate for a
$n=0.5$ polytrope). The orbital frequency at the stellar surface can be
written, to linear order in $\nu_s$, as
\begin{equation}
\nu_K(R)=2169\,M_{1.4}^{1/2}R_{10}^{-3/2}\left[1-0.094\,a\,
\left({M_{1.4}\over R_{10}}\right)\right]\,{\rm Hz}.
\end{equation}
Note that in the above equations, $R$ refers to the equatorial radius
of the (spinning) neutron star, which is related to the radius, $R_0$,
of the corresponding nonrotating star by:
\begin{equation}
{R-R_0\over R_0}\simeq 0.4{\Omega_s^2 R_0^3\over GM}
\simeq 0.0078\,M_{1.4}^{-1}R_{10}^3\left({\nu_s\over 300\,{\rm Hz}}
\right)^2,
\end{equation}
where we have again adopted the numerical parameters
appropriate for a $n=0.5$ polytrope (Lai et al.~1994).
One may appeal to numerical calculations (e.g., Miller, Lamb \& Cook
1998) for more accurate results, but the approximate expressions given above
are adequate.
Figure 2 shows the contours of constant $\nu_{\rm max}={\rm min}[\nu_K(r_{\rm
iso}),\nu_K(R)]$ in the $M$-$R_0$ plane. For large $M$ and small $R_0$,
the contours are specified by $\nu_K(r_{\rm iso})$, while for larger $R_0$
and small $M$, the contours are specified by $\nu_K(R)$.
We see that to obtain the maximum
QPO frequency of order $1100-1200$~Hz, one can either have
a $M\go 2M_\odot$ neutron star (with $R_0\lo 16$~Km), or have
a $M\simeq 1.4M_\odot$ neutron star with $R_0\simeq 14-15$~km. Here we
focus on the latter interpretation, in which
the accretion disk
terminates at the stellar surface before reaching the ISO. A
boundary layer forms in which the angular velocity of the accreting
gas changes from near the Keplerian value (at the outer edge of the
boundary layer) to the stellar rotation rate. Depending on the thickness
of the boundary layer, the inferred the NS radius may be somewhat
smaller. Moreover, the peak rotation frequency may be below $\nu_K(R)$,
which would also allow smaller values of $R_0$.
In addition to avoiding a large neutron star mass (see \S2),
the identification of $\nu_{\rm max}$ with
the Kepler frequency near the stellar surface may allow
a plausible explanation of the observed correlation between
the QPO amplitude and the X-ray flux. While the mechanism
of producing X-ray modulation in a kHz QPO is uncertain,
in many models (e.g., Miller et al.~1998; see also Klu\'zniak et al.~1990)
the existence of a supersonic ``accretion gap'' between the
stellar surface and the accretion disk is crucial for
generating the observed the X-ray modulation.
If we interperate $\nu_{\rm max}$ as
the Kepler frequency at the ISO, which is always outside the stellar
surface, then the ``accretion gap'' always exists,
and there is no qualitative change in the flow behavior as the inner disk
approaches ISO. It is therefore difficult to explain why the QPO
amplitude decreases and eventually vanishes as the X-ray flux increases.
The situation is different if $\nu_{\rm max}=\nu_K(R)$, since the gap disappears
when the mass accretion rate becomes sufficiently large.
At small $\dot M$ there is a gap (induced by a combination
of magnetic and GR effects) between the inner edge
of the disk and the stellar surface. Since the impact velocity
of the gas blob at the stellar surface is larger for a wider accretion gap,
we expect the modulation amplitude to be larger for small accretion rates
\footnote{When $\dot M$ is too low (for a given $B_0$)
so that $r_{\rm in}$ is far away from the
stellar surface, the accreting gas can be channeled out of the disk plane by
the magnetic field toward the magnetic poles. The detail
of the channeling process depends on the magnetic field geometry
in the disk (such as the radial pitch angle of the field line).
This may quench the kHz QPOs
and give rise to X-ray pulsation (as in X-ray pulsars).
The pulsating X-ray transient system SAX J1808.4-3658 may just be such an
example.}.
As $\dot M$ increases, the inner disk edge approaches the stellar surface,
and we expect the QPO amplitude to decrease. The maximum QPO
frequency signifies the closing of the accretion gap and the formation of a
boundary layer. Since there is no supersonic flow in this case,
one might expect the QPO amplitude to vanish. In addition, there may be
changes in the spectral properties of the system as the gap closes.
The large neutron star radius ($15$~km for a $1.4M_\odot$ star)
required if $\nu_{\rm max}=\nu_K(R)$ is only allowed
for a handful of very stiff
nuclear equations of state (see Fig.~2); most recent microscopic
calculations give $R_0\sim 10$~km (e.g., Wiringa et al.~1988).
Is such a large radius consistent with observations?
No neutron star radii are known with the accuracy that has been
achieved for numerous neutron star mass determinations, but several
methods have been tried:
\begin{enumerate}
\item Observations of X-ray bursts have been
used to determine empirical $M-R$ relations, but these are
hampered by the need for model-dependent assumptions regarding the total
luminosity and its time history, anisotropy of the emission, radiated
spectrum and surface composition, even when the source distance is
known (e.g. van Paradijs et al. 1990, Lewin, van Paradijs \& Taam 1995).
\item X-ray
and optical observations of the (apparently nonrotating) isolated
neutron star RX J185635-3754 (Walter, Wolk \& Neuh\"auser 1996, Walter \& Matthews
1997), combined with limits on the source distance, $D$, imply a
blackbody radius $R(1+z)<14(D/130\,{\rm pc})$ km, where $z$ is the
surface redshift of the star.
\item Ray tracing and lightbending may be used to derive limits
on $R/M$ for periodically modulated X-ray emission. For two isolated
neutron stars (PSR B1929+10 and B0950+08; Yancopoulos, Hamilton \& Helfand 1994,
Wang \& Halpern 1997) and one millisecond pulsar (J0437-4715; Zavlin
and Pavlov 1997, Pavlov \& Zavlin 1998), the results are broadly consistent with
$Rc^2/2GM\simeq 2.0-2.5$, but the results depend on geometry (angles
between rotation and magnetic axes, and rotation axis and the line of
sight) as well as on the spectrum and (energy-dependent) anisotropy
of the polar cap emission. The rather large observed pulsed fractions
appear to rule out two polar cap hot spots unless $Rc^2/2GM$ is
rather large (e.g. $\simeq 4.3$ for PSR B1929+10; Wang \& Halpern
1997). Similar considerations may prove fruitful for periodically
modulated flux from X-ray bursts (e.g. Miller \& Lamb 1998); the
pulse fractions observed so far are large, suggesting non-compact
sources (e.g. Strohmayer et al. 1999, who find $Rc^2/2GM\simeq 5$
for 4U1636-54, corresponding to an implausibly large
radius of 21 km for $M=1.4M_\odot$).
\item Burderi \& King (1998) have argued that requiring the Alfv\'en
radius to be intermediate between $R$ and the corotation radius,
$R_{co}=(GM/\Omega_s^2)^{1/3}$, for the 2.5 ms
pulsating source SAX J1808.4-3658 (discovered by Wijnands \& van der
Klis 1998)
implies an upper bound of $R<13.8(M/M_\odot)^{1/3}$ km, since the pulsations
are detected at the same frequency for X-ray count rates spanning an
order of magnitude. However, their bound
depends on the model-dependent assumptions
that the count rate is strictly proportional to $\dot M$ and the field
strength in the disk is dipolar ($B\propto r^{-3}$).
(See also Psaltis \& Chakrabarty 1999.)
\end{enumerate}
\noindent Taken together, the evidence neither supports nor excludes the
possibility that $R\simeq 15$ km for $M\simeq 1.4M_\odot$ (or $Rc^2/2GM
\simeq 3.6$) definitively, although most of the estimates listed above
favor more compact models ($R\simeq 10$ km for $M\simeq 1.4M_\odot$)
nominally.
\subsection{QPOs from $r>r_{\rm in}$?}
QPOs are identified in the Fourier spectra of photon counts from X-ray
sources, so it may be that most of the spectral power comes from radii
outside $r_{\rm in}$, possibly from the disk radius at which the differential
photon emission rate is maximum. For example, if the QPO arises from
a radius $r=(1+\lambda)r_{\rm in}$, then $\nu_{\rm QPO}=(1+\lambda)^{-3/2}
\nu_K(r_{\rm in})$. As $r_{\rm in}\to 6GM/c^2$, the ISO in the slow-rotation limit,
$\nu_{\rm QPO}\to 2200\,{\rm Hz}/(M/M_\odot)(1+\lambda)^{3/2}$, so observations that
give $\nu_{\rm max}\simeq 1060\,{\rm Hz}$ asymptotically may actually require
$M(1+\lambda)^{3/2}=2.1M_\odot$, or $1+\lambda\approx 1.3$ if
$M\approx 1.4M_\odot$. rather than $M\simeq 2.1M_\odot$.
To obtain a simple realization of this idea, consider a Shakura-Sunyaev
(1973) disk, for which the emitted flux from one face is
\begin{equation}
F(r)=\sigma_{\rm SB}T_e^4(r)={3GM\dot M f(r)\over 8\pi r^3};
\end{equation}
in the Newtonian limit (which we shall employ here for giving a
simplified illustration). The function
$f(r)=1-\beta\sqrt{r_{\rm in}/r}$, where
$\beta\leq 1$ parametrizes the rate of accretion of
angular momentum from the disk onto the star relative to
$\dot M\sqrt{GMr_{\rm in}}$
(e.g. Shapiro \& Teukolsky 1983. eq. [14.5.17]; see also
Frank, King \& Raine 1992, \S 5.3); if ``imperfect'' fluid stresses vanish
at $r_{\rm in}$, then $\beta=1$ (as in black hole accretion; see Page
\& Thorne 1974, Novikov \& Thorne 1973). If the
color temperature of the emission equals the effective temperature
$T_e(r)$, then the ``bolometric flux'' of photons is $\sim F(r)/kT_e(r)$
at radius $r$, and the rate at which photons are emitted from radii
between $r$ and $r+dr$ is of order
\begin{equation}
{2\pi rF(r)\over kT_e(r)}\sim {\dot M^{3/4}[f(r)]^{3/4}\over
r^{5/4}}.
\label{eq:noemit}
\end{equation}
Differentiating equation (\ref{eq:noemit}) implies a maximum emission
rate at $\sqrt{r/r_{\rm in}}=1.3\beta$, consistent with $r>r_{\rm in}$ provided that
$\beta>0.77$. Assuming that the QPO frequency is the Kepler frequency
at the radius of peak (bolometric) photon emission,
\begin{equation}
\nu_{\rm QPO} = {\nu_K(r_{\rm in})\over (1.3\beta)^3}
\to{1000\,{\rm Hz}\over\beta^3M/M_\odot},
\end{equation}
where the limiting result
is for $r_{\rm in}\to 6GM/c^2$. In order for the maximum
value of $\nu_{\rm QPO}$ to be $\nu_{\rm max}\simeq 1060$ Hz, we require
$M=1.4M_\odot/(\beta/0.88)^3$.
Real disk emission profiles for small $r_{\rm in}$, and the determination of
$\nu_{\rm QPO}$, are not this simple for several reasons. A detailed
calculation of the X-ray spectrum is needed, since the QPOs are found
for counts in particular energy bands; the bolometric count rate is
not a good approximation in general. (But note that Comptonization by
hot coronal gas above the disk conserves photon number, so the
approximation may be better than it appears at first sight.) In
particular, the color temperature is not usually the same as the
effective temperature, since electron scattering is the dominant
opacity at relevant disk radii. The composition of the disk is also
important; at low enough $\dot M$, the disk will be matter-dominated, but
at larger $\dot M$, radiation-dominated. (Less important, but still
significant, is the dependence of opacity on the element abundances
in the accreting gas.) In addition, relativistic effects alter
$f(r)$ (e.g. Page \& Thorne 1974, Novikov \& Thorne 1973), and hence
$\nu_{\rm QPO}$. Moreover, the angular momentum carried away by photons
may not be insignificant once $r_{\rm in}$ approaches the ISO (Page \&
Thorne 1974, Epstein 1985). Instabilities associated with the transition
from matter to radiation domination (Lightman \& Eardley 1974) or the
inner boundary layer (e.g. Epstein 1985) might also play a role in
determining $\nu_{\rm QPO}$. These and other issues associated with the
termination of disks at $r_{\rm in}$ and QPOs will be explored more fully
elsewhere. However, the simplified example presented here
indicates that $\nu_{\rm QPO}$ might plausibly arise from
$r>r_{\rm in}$.
\section{Conclusion}
In this paper we have presented a phenomenological model
of the inner region of the accretion disk for weakly magnetized
neutron stars such as those in LMXBs. A notable feature of these systems
is that both magnetic field and general relativity are important in
determining the inner disk radius. Our result suggests that the combined
effects of a steepening magnetic field -- which is likely for disk
accretion onto a neutron star --
and general relativity can produce the flattening of
the QPO frequency $\nu_{\rm QPO}$ as the mass accretion rate $\dot M$
increases. If the field steepens fast enough with decreasing inner
disk radius, $\nu_{\rm QPO}$ may vary little over a fairly substantial range
of $\dot M$ at values considerably below the
Kepler frequency at the ISO due to general relativity.
Observationally, the correlation between $\nu_{\rm QPO}$ and
the RXTE photon count rate has been well-established, but the scaling
between $\nu_{\rm QPO}$ and $\dot M$ is ambiguous (Mendez et al.~1998c).
An observational or phenomenological
determination of this scaling would be quite useful
in constraining the magnetic field structure in LMXBs.
Currently it is not clear whether the plateau behavior in
the QPO frequency has been observed. But even if $\nu_{\rm QPO}\sim 1100
-1200$~Hz represents the maximum possible QPO frequency, we argue that
a massive neutron star ($M\go 2M_\odot$) is not necessaily implied.
Instead, a $M\simeq 1.4M_\odot$, $R_0\sim 14-15$~km neutron star
may be a better solution, and is within the range allowed by
some nuclear equations of state. If this is the case,
the maximum QPO frequency signifies the closing of the accretion gap and the
formation of a boundary layer.
Alternatively, the QPO frequency might be associated with the Kepler
frequency at a radius somewhat larger than the inner radius of the disk,
thus allowing lower mass for the accreting neutron star. In either case,
better theoretical and phenomenological
understanding of the termination of magnetized accretion disks
is needed before observations of maximal kHz QPOs can be interpreted as purely
general relativistic in origin, and used to deduce neutron star masses.
\acknowledgments
D.L. is supported
by a Alfred P. Sloan Foundation Fellowship.
R.L. acknowledges support from NASA grant NAG 5-6311.
I.W. acknowledges support from NASA grants NAG 5-3097 and NAG 5-2762.
|
\section{Introduction}
The aim of this article is to study the generalization of the annihilator
theorem of Duflo to
a certain class of Lie superalgebras, namely to the class of orthosymplectic
Lie superalgebras $\osp(1,2l)$, $l\geq 1$. The main reason for considering
this class of Lie superalgebras is the complete reducibility of finite
dimensional modules.
Both in classical and quantum cases, the complete reducibility
appears in various steps of the study of annihilators of
Verma modules. A Lie superalgebra whose finite dimensional modules are
completely reducible is called completely reducible.
According to the theorem of
Djokovi\'c and Hochschild (see~\cite{sch}, p. 239), any finite dimensional
completely reducible Lie superalgebra is a direct sum of
semisimple Lie algebras
and algebras $\osp(1,2l)$ ($l\geq 1$).
Another useful property of $\osp (1,2l)$ is that its enveloping
algebra is a domain (see~\cite{al}).
Let ${\frak g}={\frak g}_0\oplus {\frak g}_1$ be a Lie superalgebra
$\osp(1,2l)$, and ${\frak h}$ be a Cartan
subalgebra of the Lie algebra ${\frak g}_0\simeq \ssp(2l)$. Let
$\Delta=\Delta_0\cup \Delta_1$ be the set of roots of ${\frak g}$
with respect to ${\frak h}$, where
$\Delta_0$ (resp., $\Delta_1$) is the set of even (resp., odd) roots.
Let $\pi$ be a set of simple roots of $\Delta$ and $P^+(\pi)$ be the
set of integral dominant weights.
Denote by $\Delta^+$ (resp., $\Delta_0^+,\ \Delta_1^+$)
the set of positive roots (resp., positive even, positive odd).
Set $\overline{\Delta_0^+}\!=\!\Delta_0^+\setminus2\Delta_1^+,\
\rho\!=\!{1\over 2}\mathop{\sum}_{\alpha\in \Delta_0^+}\alpha-
{1\over 2}\mathop{\sum}_{\alpha\in \Delta_1^+}\alpha$. Denote
by ${\cal U}({\frak g})$ the universal enveloping algebra
of ${\frak g}$, by ${\cal Z}({\frak g})$ its centre. Let
${\cal F}$ be the canonical filtration of ${\cal U}({\frak g})$.
Consider the case $l=1$, i.e. ${\frak g}=\osp(1,2)$, which has been treated by
Pinczon.
In this case, any ${\frak g}$-Verma module ${\widetilde M}$,
viewed as a ${\frak g}_0$-module, is the direct sum of two
${\frak g}_0$-Verma modules $M_0$ and $M_1$. Let $C_0$ be
a Casimir element for ${\frak g}_0$.
Then $C_0$ acts by
scalars $c_i$ on $M_i$ (i=0,1). In general, $c_1\not=c_0$,
and in this case, Pinczon proved that
$\Ann {\widetilde M}={\cal U}({\frak g})\Ann_{{\cal Z}({\frak g})}
{\widetilde M}$. However when the highest weight of $\widetilde{M}$ equals
$-\rho$, one has $c_1=c_0$ and so $C_0-c_0$ belongs to the annihilator of
$\widetilde{M}$. It is easy to check that $(C_0-c_0)\not\in {\cal U}({\frak g})
\Ann_{{\cal Z}({\frak g})}\widetilde{M}$.
Consequently, the annilihilator theorem
doesn't hold in full generality.
Return to the general case ${\frak g}=\osp(1,2l)$. For any
$\lambda\in {\frak h}^*$, let $\widetilde{M}(\lambda)$ be the Verma
module of highest weight $\lambda$.
Musson showed (see~\cite{mu},5.6)
that if $(\lambda+\rho,\beta)=0$ for some $\beta\in\Delta_1^+$ and
$(\lambda+\rho,\alpha)\not=0$ for all $\alpha\in \overline{\Delta_0^+}$
then $\Ann_{{\cal U}({\frak g})}\widetilde{M}(\lambda)$ strictly contains
${\cal U}({\frak g})\Ann_{{\cal Z}({\frak g})}\widetilde{M}(\lambda)$
and so the annihilator theorem does not hold.
At the end of the article~\cite{mu}, Musson
asked for which $\lambda\in {\frak h}^*$ the annihilator of
$\widetilde{M}(\lambda)$
is generated by its intersection with the centre.
In this article we give an answer to this question by proving the
following theorem:
\begin{thm}{} For any $\lambda\in {\frak h}^*$ one has
\begin{equation}
\label{thmintr}
\Ann_{{\cal U}({\frak g})}\widetilde{M}(\lambda)=
{\cal U}({\frak g})\Ann_{{\cal Z}({\frak g})}{\widetilde M}(\lambda)\ \
\Longleftrightarrow \ \ \forall\alpha\in \Delta_1^+:\
(\lambda+\rho,\alpha)\not=0.
\end{equation}
\end{thm}
Our strategy to prove this theorem is based on
a recent alternative proof (see~\cite{jfr}) of
the classical annihilator theorem which follows essentially the
quantum case treated by Joseph and Letzter--- see~\cite{jl}.
This proof has the following main steps.
1) The first ingredient is the separation theorem established
by Musson in~\cite{mu}, 1.4. Using complete reducibility,
he proved the existence of an $\ad{\frak g}$-submodule
${\cal H}$ of
${\cal U}({\frak g})$ such that the multiplication map induces an isomorphism
${\cal U}({\frak g})\simeq {\cal Z}({\frak g})\otimes {\cal H}$ of
$\ad{\frak g}$-modules. Moreover, the multiplicity of any finite dimensional
module in ${\cal H}$ is equal to the dimension of its zero weight space.
Since the centre acts on any Verma module $\widetilde{M}(\mu)$
as a scalar, the separation theorem implies that
$\Ann_{{\cal U}({\frak g})}\widetilde{M}(\mu)=
{\cal U}({\frak g})\Ann_{{\cal Z}({\frak g})}\widetilde{M}(\mu)$ iff
$\Ann_{{\cal H}}\widetilde{M}(\mu)=0$.
2) In Section~\ref{hesselink} we obtain a
formula giving the multiplicity of every simple finite dimensional
module $\widetilde{V}(\lambda)$,
$\lambda\in P^+(\pi)$, in the graded components $H^n$ where
$H=\gr_{\cal F}{\cal H}$.
This is an analogue of the classical Hesselink formula (see~\cite{he}).
However, in our case the formula looks rather suprising.
Namely, for every $\lambda\in P^+(\pi)$, the multiplicity of every
$\widetilde{V}(\lambda)$ in ${\cal H}$ is given by the formula
\begin{equation}
[H^n:\widetilde {V}(\lambda)]=\sum_{w\in W} (-1)^{l(w)} P_n(w.\lambda),
\label{superHss}
\end{equation}
where $P_n(\mu)$ are integers defined by generating function
\begin{equation}
\prod_{\alpha\in \Delta^+_0}(1-q e^{\alpha})^{-1}
(1+q^{2l}e^{\beta_1})\prod_{\beta\in {\Delta^+_1\setminus\{\beta_1\}}}
(1+e^{\beta})=
\sum_{r=0}^{\infty}\sum_{\nu\in {\Bbb N}\pi}P_r(\nu) e^{\nu} q^r
\label{superp}
\end{equation}
where $\beta_1$ is the highest weight of the ${\frak g}_0$-module
${\frak g}_1$.
3) Any Verma module $\widetilde{M}(\mu)$ has a simple
Verma submodule $\widetilde{M}(\mu')$ and $\Ann_{{\cal Z}({\frak g})}
\widetilde{M}(\mu)
=\Ann_{{\cal Z}({\frak g})}\widetilde{M}(\mu')$. Moreover,
$(\mu+\rho,\alpha)=0$ for some $\alpha\in \Delta_1^+$ iff
$(\mu'+\rho,\alpha')=0$ for some $\alpha'\in \Delta_1^+$. Thus, in order to
prove the implication "$\Leftarrow$" of~(\ref{thmintr}), it is sufficient
to check it for simple Verma modules.
For the other implication, we show (see~\ref{otherim}) that there
exists an $\ad {\frak g}$ submodule $V$ of ${\cal U}({\frak g})$, which
lies in the annihilator of any simple Verma module
$\widetilde{M}(\mu)$ such that $(\mu+\rho,\alpha)=0$ for some $\alpha\in
\Delta_1^+$.
Then,~\Lem{zarcl} implies that $V$ lies indeed in the annihilator
of any Verma module $\widetilde{M}(\mu)$ such that $(\mu+\rho,\alpha)=0$
for some $\alpha\in \Delta_1^+$.
So in a sense, both implications of the equivalence~(\ref{thmintr}) are reduced
to the case of a simple Verma module.
4) In Section~\ref{shapo} we present a criterion of simplicity for a
Verma module
$\widetilde{M}(\mu)$ given by Kac (see~\cite{k3} and also~\cite{mu2} 2.4).
As in the classical case, this criterion is related to the so-called
Shapovalov determinants $\det { S}_{\nu}\in
{\cal U}({\frak h})$, $\nu\in {\Bbb N}\pi$. These determinants are constructed
in such a way that $\det {S}_{\nu}(\mu)=0$
means that the Verma module $\widetilde{M}(\mu)$ contains a proper submodule
with a non trivial element of weight $\mu-\nu$. Thus, $\widetilde{M}(\mu)$
is simple iff $\det S_{\nu}(\mu)\not=0$ for all $\nu\in{\Bbb N}\pi$.
Kac proved
that all factors of Shapovalov determinants are linear and he gave an explicit
formula for the
factorization of these determinants (see~\ref{shapovalov}).
5) The separation theorem allows one to define
the Parthasarathy--Ranga-Rao--Varadarajan (PRV) determinants as in the
classical case (see~\ref{prvcons} for details).
For any $\lambda\in {\frak h}^*$, denote by $\widetilde{V}(\lambda)$
the unique simple module of the highest weight
$\lambda$.
The PRV determinant $\det{PRV}^{\lambda}\in {\cal U}({\frak h})$,
$\lambda\in P^+(\pi)$, has the property:
$$\forall \mu\in {\frak h}^*,\
\det {PRV}^{\lambda}(\mu)=0\Longleftrightarrow
\exists \widetilde{V}(\lambda)\subset {\cal H},
\widetilde{V}(\lambda)\subset \Ann_{\cal H}\widetilde{V}(\mu).$$
Thus $\Ann_{{\cal H}}\widetilde{V}(\mu)=0$ iff
$\det{PRV}^{\lambda}(\mu)\not=0$ for all $\lambda\in P^+(\pi).$
Therefore we have to prove that for any simple Verma module
$\widetilde{M}(\mu)=\widetilde{V}(\mu)$ one has the following equivalence
\begin{equation}
\label{prvshap}
\exists \lambda\in P^+(\pi):\ \det{PRV}^{\lambda}(\mu)=0\ \ \Longleftrightarrow
\ \ \exists\alpha\in \Delta_1^+:\ (\mu+\rho,\alpha)=0.
\end{equation}
In Section~\ref{prv} we describe the zeroes of the determinants
$\det{PRV}^{\lambda}$.
Both in the classical semisimple and quantum cases
the union of the zeroes of $\det{PRV}^{\lambda}$ ($\lambda\in P^+(\pi)$)
coincides with the union of zeroes of the Shapovalov determinants
$\det { S}_{\nu} (\nu\in {\Bbb N}\pi)$.
This implies the annihilation theorem.
6) In~\Thm{thmprv} we give a linear factorization
of the determinants $\det{PRV}^{\lambda}:\ \lambda\in P^+(\pi)$.
There are factors of two types.
The factors of the first type, called "standard factors", coincide
with factors of Shapovalov determinants. They
can be obtained as reviewed in ~\cite{jfr}. We
briefly summarize this procedure in~\ref{stfactor}---~\ref{pfthmLL}.
These factors have form
\begin{equation}
\label{stnd}
\begin{array}{lcc}\varphi(\alpha)+(\rho,\alpha)-m(\alpha,\alpha)/2, &\alpha\in
\overline{\Delta_0^+},& m\in {\Bbb N}, m\geq 1\\
\varphi(\alpha)+(\rho,\alpha)-(2m+1)(\alpha,\alpha)/2, &\alpha\in\Delta_1^+, &
m\in {\Bbb N}
\end{array}
\end{equation}
where the element $\varphi(\alpha)$ of ${\frak h}$ is defined by the formula
$\varphi(\alpha)(\mu)=(\alpha,\mu)$.
A delicate point is to get the factors of the second type, called
"exotic factors". They have form
\begin{equation}
\label{intrex}
\varphi(\alpha)+(\rho,\alpha),\ \alpha\in \Delta_1^+.
\end{equation}
The Hesselink formula discussed in 2) ensures that factors (\ref{stnd}),
(\ref{intrex}) indeed exhaust
the set of factors of PRV determinants.
The exotics factors
correspond precisely to the equivalence~(\ref{prvshap}).
We conclude that the annihilator of a simple Verma
module
$\widetilde{M}(\mu)=\widetilde{V}(\mu)$ is generated by its intersection
with the centre ${\cal Z}({\frak g})$ iff $\mu$ is not a zero of an
exotic factors of some PRV determinant.
\begin{rem}{} Taking into account that
any completely reducible Lie superalgebra is a direct sum
of simple Lie algebras and Lie superalgebras of the type
$\osp(1,2l)$, our results (theorems~\ref{thmhess},~\ref{thmprv},~\ref{main})
can be directly extended to the case of any
finite dimensional completely reducible Lie
superalgebra ${\frak g}$.
\end{rem}
{\em Acknowledgement.} We are greatly indebted to our teacher
Prof. A.~Joseph who proposed the problem and pushed us to solve it.
His paper "Sur l'annulateur d'un module de Verma" was the
main inspiration of the present work.
We would like to thank V.~Hinich for numerous suggestions and support.
We are also grateful to T.~Levasseur and M.~Duflo for helpful discussions
and important remarks.
\section{background}
\label{notation}
In this Section we fix the main notations we use throughout this paper.
The notation ${\Bbb N}^+$ will stand for the set of positive integers.
The base field we are going to work with is ${\Bbb C}$.
\subsection{}
Let ${\frak g}$ be the Lie superalgebra $\osp(1,2l),\ l\geq 1$
(see Kac~\cite{k2} for a
presentation of this Lie superalgebra by generators and relations). Denote
by ${\frak g}_0$ the even part and by ${\frak g}_1$
the odd part of ${\frak g}$.
We recall that ${\frak g}_0\simeq \mbox{sp}(2l)$. Fix a Cartan subalgebra
${\frak h}$ in ${\frak g}_0$. Denote by $\Delta_0$ (resp., $\Delta_1$)
the set of even (resp., odd) roots of ${\frak g}$.
Set $\Delta=\Delta_0\cup\Delta_1$.
Let $\Delta_{irr}$ be the set of irreducible roots
of $\Delta$. Then $\Delta_{irr}=\overline{\Delta_0}\cup \Delta_1$, where
$\overline{\Delta_0}\!\!:=\!\Delta_0\backslash 2\Delta_1$.
Fix a basis of simple roots $\pi$ of $\Delta$, and define correspondingly
the sets
$\Delta^{\pm}, \Delta_0^{\pm},\Delta_1^{\pm},\overline{\Delta_0}^{\pm},
\Delta_{irr}^{\pm}$. Denote by $W$ the Weyl group of $\Delta$.
Set
$$\begin{array}{ccc}
\rho_0\!:={1\over 2}\displaystyle\mathop{\sum}_{\alpha\in \Delta_0^+}\alpha,&
\rho_1\!:={1\over 2}\displaystyle\mathop{\sum}_{\alpha\in \Delta_1^+}\alpha,&
\rho\!:=\rho_0-\rho_1={1\over 2}\displaystyle\mathop{\sum}_{\alpha\in
\Delta_{irr}^+}\alpha.
\end{array}$$
Introduce the standard partial order relation on ${\frak h}^*$:
$\lambda\leq \mu \Longleftrightarrow \mu-\lambda\in {\Bbb N}\pi$.
Denote by $(-,-)$ the non-degenerate bilinear form on ${\frak h}^*$ coming
from the
restriction of the Killing form of ${\frak g}_0$
to ${\frak h}$. Let $\varphi:{\frak h}^*\longrightarrow {\frak h}$ be the
isomorphism
given by $\varphi(\lambda)(\mu):=(\lambda,\mu)$. For any $\lambda,\mu\in
{\frak h}^*,\
\mu\not=0$ one defines
$$\langle\lambda,\mu\rangle:=2\displaystyle{(\lambda,\mu)\over (\mu,\mu)}.$$
\subsection{}
\label{realization}
One has the following useful realization of $\Delta$.
Identify ${\frak h}^*$ with ${\Bbb C}^l$ and consider $(-,-)$ as
a scalar product on ${\Bbb C}^l$. Then there exists an orthonormal basis
$\{\beta_1,\ldots,\beta_l\}$ such that
$$\begin{array}{lr}\pi=\{\beta_1-\beta_2,\ldots,\beta_{l-1}-\beta_l,\beta_l\}&
\\
\Delta_0^+=\{\beta_i\pm\beta_j, 1\leq i<j\leq l,\ 2\beta_i,\ 1\leq i\leq l\},&
\Delta_1^+=\{\beta_i, 1\leq i\leq l\}\\
\Delta_{irr}^+=\{\beta_i\pm\beta_j,\ \beta_i,\ 1\leq i<j\leq l\},&
\overline{\Delta_0}^+=\{\beta_i\pm\beta_j,\ 1\leq i<j\leq l\}\\
\rho=\displaystyle\sum_{i=1}^l(l-i+{1\over 2})\beta_i,&
\rho_0=\displaystyle\mathop{\sum}_{i=1}^l(l-i+1)\beta_i
\end{array}$$
and the Weyl group $W$ is just the group of the signed permutations of the
$\beta_i$.
Define the translated action of $W$ on ${\frak h}^*$ by the formula:
$w.\lambda=w(\lambda+\rho)-\rho$ for $\lambda\in {\frak h}^*$ and $w\in W$.
For any element $w\in W$ set $\sgn w:={(-1)}^{l(w)}$, where
$l(w)$ is the length of a reduced decomposition of $w$.
There exists a Chevalley antiautomorphism $\sigma$ of ${\frak g}$
of order $4$ which leaves invariant the elements of ${\frak h}$ and
maps ${\frak g}_{\alpha}$ to ${\frak g}_{-\alpha}$ for any $\alpha\in \Delta$.
\subsection{Enveloping algebra}
As usual, if ${\frak k}$ is a Lie superalgebra,
${\cal U}({\frak k})$ denotes its enveloping algebra.
Set ${\cal F}$ the natural
filtration of ${\cal U}({\frak g})$ defined by
${\cal F}={({\frak g}^n)}_{n\in \Bbb N}$.
The graded algebra of ${\cal U}({\frak g})$ associated to ${\cal F}$ is
the symmetric superalgebra denoted by
${\cal S}({\frak g})\simeq {\cal S}({\frak g}_0)\otimes\bigwedge{\frak g}_1$
which is not a domain. Nevertheless,
Aubry and Lemaire showed (\cite{al}) that ${\cal U}({\frak g})$ is a domain.
We define the supercentre to be the vector subspace of ${\cal U}({\frak g})$
generated by the homogenous elements $a$ such that $ax={(-1)}^{|a||x|}xa$
for all homogenous elements $x$ in ${\cal U}({\frak g})$.
For ${\frak g}=\osp (1,2l)$, the supercentre coincides with the
genuine centre.
The Lie superalgebra ${\frak g}$ acts on ${\cal U}({\frak g})$ and
${\cal S}({\frak g})$ by superderivation
via the adjoint action. We denote these actions by $\ad$.
Throughout this paper,
an action of any element of ${\frak g}$ on ${\cal U}({\frak g})$ means always
the adjoint action.
We identify ${\cal U}({\frak h})$ with ${\cal S}({\frak h})$.
\subsection{Verma and simple highest weight modules}
For any $\alpha\in \Delta$, let ${\frak g}_{\alpha}$ be the one-dimensional
subspace of weight $\alpha$ of ${\frak g}$ and
${\frak n}^{\pm}=\displaystyle\mathop{\bigoplus}_{\alpha\in \Delta^{\pm}}
{\frak g}_{\alpha}$,\ ${\frak b}^{\pm}=\frak h\oplus{\frak n}^{\pm}$.
For a fixed ${\lambda}\in {\frak h}^*$, let $\widetilde{\Bbb C}_{\lambda}$ be
the one dimensional ${\frak b}^+$-module with ${\frak n}^+v=0$ and
$hv=\lambda(h)v$ for all $h\in {\frak h}$ and
$v\in \widetilde{\Bbb C}_{\lambda}$. Set
$$\widetilde{M}(\lambda)={\cal U}({\frak g}){\otimes}_{{\cal U}({\frak b}^+)}
\widetilde{\Bbb C}_{\lambda}.$$
The Verma module $\widetilde{M}(\lambda)$ has a unique simple quotient
denoted by
$\widetilde{V}(\lambda)$.
\subsection{The $\widetilde{\cal O}$ category}
\label{ocat}
Let $M$ be a ${\frak g}$-module. For any $\lambda\in{\frak h}^*$, set
$$M_{\lambda}=\left\{m\in M|\ hm=\lambda(h)m,\
\forall h\in {\frak h}\right\}.$$
A non-zero vector $v\in M$ has weight $\lambda$ if $v\in M_{\lambda}$.
For any subspace
$N$ of $M$ we denote by $\Omega(N)$ the set of weights $\lambda\in{\frak h}^*$
such
that $N\cap M_{\lambda}\not=\{0\}$. The module $M$ is said to be
diagonalizable if
$M=\displaystyle\mathop{\bigoplus}_{\lambda\in{\frak h}^*}M_{\lambda}$.
If $M$ is a diagonalizable module and
$\dim M_{\lambda}<\infty$ for all $\lambda\in {\frak h}^*$, we set
$\ch M=\displaystyle\mathop{\sum}_{\lambda\in{\frak h}^*}
\dim M_{\lambda}e^{\lambda}$.
Throughout this article, we shall work essentially with two categories of
${\frak g}$-modules.
The first one is the category of completely reducible ${\frak g}$-modules.
If $M$ is such a module and
$V$ is any simple module, we shall denote by $[M:V]$ the number
$\dim \Hom_{\frak g} (V,M)$.
The second one, denoted by $\widetilde{\cal O}$, is the full
subcategory of the category of
${\frak g}$-modules whose objects are ${\frak g}$-modules $M$ such that
\begin{enumerate}
\item $M$ is diagonalizable
\item $\forall \lambda\in \Omega(M),\dim M_{\lambda}<+\infty$
\item $\exists \lambda_1,\ldots,\lambda_k\in {\frak h}^* |\
\Omega(M)\subset \mathop{\bigcup}_{i=1,\ldots, k}(\lambda_i-{\Bbb N}\pi)$
\end{enumerate}
Any object of the $\widetilde{\cal O}$ category, considered as
a ${\frak g}_0$-module, belongs to the ${\cal O}$ category
(see~\cite{jfr}, 3.5 for definition).
In particular, any ${\frak g}$-module $M$ of $\widetilde{\cal O}$
has finite length. If $V$ is a simple highest weight module, we denote
by $[M:V]$ the number of simple quotients isomorphic to $V$ in any
composition series of $M$. This number does not depend of the choice of the
composition series of $M$.
\subsection{Finite dimensional representations}
\label{findim}
Define for $r\in\{1,\ldots,l\}$ the fundamental weight
$\omega_r=\displaystyle\sum_{i=1}^r\beta_i$, and introduce the set
$$P^+(\pi)\!:=\displaystyle\sum_{r=1}^l{\Bbb N}\omega_r=
\{\lambda\in {\frak h}^*|\ \langle\lambda,\beta_l\rangle\in 2{\Bbb N},\
\langle\lambda,\beta_i-\beta_{i+1}\rangle\in {\Bbb N},\
\forall i=1,\ldots,l-1\}.$$
Kac (see~\cite{k1}) showed that $\widetilde{V}(\lambda)$ is
finite dimensional iff
$\lambda\in P^+(\pi)$. For any $\mu\in{\frak h}^*$, let $V(\mu)$ be the simple
${\frak g}_0$-module of highest weight $\mu$. Remark that
$\{\beta_1-\beta_2,\ldots,\beta_{l-1}-\beta_l,2\beta_l\}$ is a basis of
simple roots of $\Delta_0$ and that $\langle \mu,2\beta_l\rangle={1\over
2}\langle\mu,\beta_l\rangle$ for all $\mu\in {\frak h}^*$.
Thus $V(\lambda)$ is finite dimensional iff $\lambda\in P^+(\pi)$.
\subsubsection{}
\label{duality} Observe that the longest element of the Weyl group acts
on ${\frak h}^*$ as $-id$. Therefore for any finite dimensional module $V$,
$\Omega(V)=-\Omega(V)$.
\section{The separation theorem}
\label{linkann}
Recall the separation theorem established by Musson in~\cite{mu},1.4:
\subsection{}
\begin{thm}{separation} There exists an ad-invariant subspace ${\cal H}$ in
${\cal U}({\frak g})$ such that the multiplication map induces an
$\ad {\frak g}$-isomorphism ${\cal U}({\frak g})\simeq
{\cal Z}({\frak g})\otimes {\cal H}$. Moreover, for every simple
finite dimensional module $\widetilde V$, $[{\cal H}:\widetilde{V}]=\dim
\widetilde{V}_0$.
\end{thm}
\subsection{}
\label{redtosimple}
Since the centre ${\cal Z}({\frak g})$ acts by a scalar on every Verma module
$\widetilde{M}(\mu)$, the separation theorem implies
the following equivalence :
$$Ann_{{\cal U}({\frak g})}\widetilde{M}(\mu)={\cal U}({\frak g})
\Ann_{{\cal Z}({\frak g})}\widetilde{M}(\mu)\ \Longleftrightarrow \
\Ann_{\cal H}\widetilde{M}(\mu)=\{0\}.$$
As in the classical case, any Verma module contains a simple Verma submodule.
Let $\widetilde{M}(\mu)$ be a Verma module and $M$ be its simple
Verma submodule. Obviously, $$\Ann_{{\cal H}}\widetilde{M}(\mu)\subset
\Ann_{{\cal H}} M.$$
Since $\widetilde{M}(\mu)$ and $M$ have the same central character,
$M$ has the form $M\simeq M(w.\mu)$ for some $w\in W$ (see~\cite{mu},1.1).
By definition of the
translated action of $W$, $w.\mu+\rho=w(\mu+\rho)$ and as $W$ acts on
$\Delta_1$ by signed permutations, one has
$$\exists \alpha\in \Delta_1^+,\ (\mu+\rho,\alpha)=0 \Longleftrightarrow
\exists \alpha\in \Delta_1^+,\ (w.\mu+\rho,\alpha)=0.$$
Hence the proof of the implication "$\Leftarrow$" of our
main~\Thm{main}
can be reduced to the case of simple Verma modules. Indeed for this,
it is enough to
show that
\begin{equation}
\label{eqiv1}
\bigr[\widetilde{M}(\mu)\mbox{ simple and } (\mu+\rho,\alpha)\not=0\ \forall
\alpha\in \Delta_1^+\bigr] \ \Longrightarrow \ \ \Ann_{{\cal H}}\widetilde{M}
(\mu)=0.
\end{equation}
\section{Factorization of the Shapovalov determinants}
\label{shapo}
\subsection{}
\label{contrform}
The triangular decomposition ${\cal U}({\frak g})={\cal U}({\frak n}^-)\otimes
{\cal U}({\frak h})\otimes
{\cal U}({\frak n}^+)$ leads to the following decomposition
$${\cal U}({\frak g})={\cal U}({\frak h})\oplus({\frak n}^-{\cal U}({\frak g})+
{\cal U}({\frak g}){\frak n}^+)$$
The projection $\Upsilon:{\cal U}({\frak g})\longrightarrow
{\cal U}({\frak h})$
with respect to this decomposition is called the
Harish-Chandra projection. Observe that $\Upsilon(ha)=h\Upsilon(a)$ for
all $h\in {\cal U}({\frak h}),a\in {\cal U}({\frak g})$.
The Harish-Chandra projection map allows us to define a contravariant form on
each Verma
module $\widetilde{M}(\lambda)$ by
the formula $${\langle av_{\lambda},bv_{\lambda}\rangle}_{\lambda}=
\Upsilon(\sigma(a)b)
(\lambda)\quad \forall a,b\in {\cal U}({\frak g}),$$
where $v_{\lambda}$ is a primitive vector of highest weight $\lambda$.
The kernel of this form is the largest proper submodule of $\widetilde{M}
(\lambda)$.
Consequently, the Verma
module $\widetilde{M}(\lambda)$ is simple if and only if the form
${\langle-,-\rangle}_{\lambda}$ is non-degenerate.
The subspaces of different weights are pairwise orthogonal with respect to
${\langle-,-\rangle}_{\lambda}$. Therefore, the quest
for a criterion of simplicity for a Verma module leads naturally to the
analysis of the zeroes of the
so called Shapovalov determinants: $\det S_{\nu}$, where $\nu\in {\Bbb N}\pi$
and
$S_{\nu}:\ {{\cal U}({\frak n}_-)}_{\nu}\times {{\cal U}({\frak n}_-)}_{\nu}
\mapsto {\cal S}({\frak h})$ is
defined by the formula $$S_{\nu}(x,y)=\Upsilon(\sigma(x)y).$$
The factorization of these determinants was established by Kac
(see~\cite{k3}) for all classical simple Lie superalgebras.
For the present case ${\frak g}=\osp(1,2l)$ (see also~\cite{mu2} 2.4)
one has the
\subsection{}
\begin{thm}{shapovalov} For all $\nu\in {\Bbb N}$, one has, up to a non-zero
scalar,
$\vtop{\hbox to \textwidth{$\det {S}_{\nu}=\displaystyle
\mathop{\prod}_{\alpha\in\overline{\Delta_0^+}}\mathop{\prod}_{m=1}^{\infty}
{(\varphi(\alpha)+(\rho,\alpha)-{1\over 2}m(\alpha,\alpha))}^
{\tau(\nu-m\alpha)}
\times$\hfill}
\hbox to \textwidth{\hfill$\displaystyle\mathop{\prod}_{\alpha\in{\Delta_1^+}}
\mathop{\prod}_{m=1}^{\infty}
{(\varphi(\alpha)+(\rho,\alpha)-{1\over 2}(2m-1)(\alpha,\alpha))}^
{\tau(\nu-(2m-1)\alpha)}$}}$
where $\tau:{\Bbb Z}\pi\longrightarrow {\Bbb N}$ is the Kostant
partition function defined by $\tau({\Bbb Z}\pi\backslash {\Bbb N}\pi)=0$
and
$$\tau(\nu)=\# \left\{ {\{k_{\alpha}\}}_{\alpha\in \Delta^+} |\
k_{\alpha}\in {\Bbb N}\mbox{ and }
\sum_{\alpha\in\Delta^+}k_{\alpha}\alpha=\nu\right\}, \ \ \forall
\nu\in{\Bbb N}\pi.$$
\end{thm}
\subsection{}
\label{weightsub}
Take $\lambda\in {\frak h}^*$.
The theorem above allows us to describe the weights of the largest proper
submodule $N$ of the Verma module $\widetilde{M}(\lambda)$:
$$\begin{array}{c}
\Omega (N)=\displaystyle\mathop{\bigcup}_{\alpha\in \Delta_{\lambda}}
\{\lambda-\langle\lambda+\rho,\alpha\rangle\alpha-{\Bbb N}\pi\},
\text { where } \\
\Delta_{\lambda}\:\!=\{\alpha\in \overline{\Delta^+_0}|\
\langle\lambda+\rho,\alpha\rangle\in {\Bbb N}^+\}\cup \{\beta\in \Delta^+_1|\
\langle\lambda+\rho,\beta\rangle\in 2{\Bbb N}+1\}.
\end{array}$$
In particular, it implies the following criterion of simplicity for
$\widetilde{M}(\lambda)$:
\subsection{}
\begin{cor}{simpleV}
The Verma module $\widetilde{M}(\lambda):\ \lambda\in {\frak h}^*$ is
simple if and only if :
$$\begin{array}{lr} \langle\lambda+\rho,\alpha\rangle\not\in{\Bbb N}^+,&
\forall \alpha\in \overline{\Delta_0^+}\\
\langle\lambda+\rho,\alpha\rangle\not\in(2{\Bbb N}+1),& \forall \alpha\in
{\Delta_1^+}.
\end{array}$$
\end{cor}
\section{Hesselink Formula}
\label{hesselink}
The goal of this section is to give a formula describing a multiplicity
of every simple finite dimensional module $\widetilde {V}(\lambda)$
in the graded component $H^n$ where $H:=\gr_{\cal F}{\cal H}$.
More precisely, let ${\cal S}^n({\frak g})
\subseteq {\cal S}({\frak g})$ be the subspace
of homogeneous elements of degree $n$. For a finite dimensional module
$\widetilde {V}(\lambda)$ we will give an explicit formula for
a Poincar\'e series $P_{\lambda}(q)$ of $H=\gr_{\cal F} {\cal H}$
which is defined by
$$P_{\lambda}(q)\!:=\sum_{n=0}^{\infty} [H^n:{\widetilde V}(\lambda)]q^n,
\text { where } H^n\!:={\cal S}^n({\frak g })\cap H.$$
\subsection{Notation}
\label{nthess}
Consider the group ring ${\Bbb Z}[{\frak h}^*]$. Its elements
are finite sums $\sum_{\lambda\in {\frak h}^*} c_{\lambda}e^{\lambda}, \
c_{\lambda}\in {\Bbb Z}$.
Our characters lie in the power series ring ${\Bbb Z}[{\frak h}^*][[q]]$.
For each $\mu\in{\frak h}^*$ define the ${\Bbb Z}[[q]]$-linear
homomorphism $\pi_{\mu}$ by
$$\pi_{\mu}: {\Bbb Z}[{\frak h}^*][[q]]\ \rightarrow\ {\Bbb Z}[[q]],\ \ \ \
\ \ e^{\lambda}\mapsto \delta_{\lambda, \mu},
\ \forall \lambda\in {\frak h}^*.$$
\subsubsection{}
\label{jprop}
For each $w\in W$ define the automorphism of the ring
${\Bbb Z}[{\frak h}^*][[q]]$ by
$$w(e^{\lambda})\!:=e^{w\lambda},\ w(q)\!:=q.$$
Let $J$ be the ${\Bbb Z}[[q]]$-linear endomorphism of
${\Bbb Z}[{\frak h}^*][[q]]$ given by the formula
$$J\!:=\sum_{w\in W} \sgn (w)w.$$
We use the following properties of the operator $J$:
\begin{enumerate}
\item $J(wa)=\sgn (w)J(a)$ for all $w\in W$. Consequently, if there exists $w$
such that $wa=a$ and $\sgn (w)=-1$ then $J(a)=0$.
\item Since for any $\lambda\in P^+(\pi)$ the stabilizer
$\Stab_W\lambda$ is generated by the simple reflections it
contains~(\cite{jq}, A.1.1) and $P(\pi)=WP^+(\pi)$, one
can conclude from (i) that
$$\Stab_W\mu\not=\{\id\}\ \Longrightarrow J(e^{\mu})=0$$
for any $\mu\in P(\pi)$.
\item Fix a subgroup $K$ of $W$ and representatives
$g_1,\ldots, g_r$ of the left cosets $W/K$. Set
$$J^{K}\!:=\sum_{w\in K} \sgn(w) w,\ \ \ \
J^{W/K}\!:=\sum_{i=1}^r \sgn(g_i){g_i}.$$
Then $J^{W/K}J^{K}=J$.
\item Assume that $wa=a$ for all $w\in K$. Then $J^K(ab)=aJ^K(b)$.
\end{enumerate}
\subsubsection{}
\label{sprth}
For a graded ${\frak h}$-submodule $N$ of ${\cal S}({\frak g})$
its graded character $\ch_q N$ is defined to be the element
of ${\Bbb Z}[{\frak h}^*][[q]]$ given by the formula
$$\ch_q N\!:=\sum_{r=0}^{\infty} \ch (N\cap {\cal S}^r({\frak g})) q^r.$$
The following relations hold
$$\begin{array}{l}
\ch_q\Lambda {\frak g}_1=\displaystyle\sum_{r=0}^{2l}
(\ch\Lambda^r {\frak g}_1)q^r=\displaystyle\prod_{\beta\in\Delta_1}
(1+qe^{\beta})\\
\ch_q{\cal S}({\frak n}_0^{\pm})=\displaystyle\prod_{\alpha\in\Delta_0^+}
(1-qe^{\pm\alpha})^{-1},\\
\ch_q{\cal S}({\frak n}^{\pm})=\displaystyle\prod_{\alpha\in\Delta^+_0}
(1-qe^{\pm\alpha})^{-1}\!\displaystyle\prod_{\beta\in\Delta^+_1}
(1+qe^{\pm\beta}),\\
\ch_q{\cal S}({\frak n}^+\oplus {\frak n}^-)=
\ch_q{\cal S}({\frak n}_0^+\oplus {\frak n}_0^-)\cdot
\ch_q\Lambda {\frak g}_1.
\end{array}
$$
Moreover, the separation theorem implies that
$$\ch_q {\cal S}({\frak g})=\ch_q {{\cal S}({\frak g})}^{{\frak g}}\ch_q H.$$
By~\cite{mu}, 1.2 ${{\cal S}({\frak g})}^{{\frak g}}\cong
{{\cal S}({\frak h})}^W$.
By~\cite{jfr}, 8.7
$$\ch_q {\cal S}({\frak h})=\ch_q {\cal S}({\frak h})^W\!\displaystyle
\sum_{w\in W}q^{l(w)}$$
so
$$\ch_q {\cal S}({\frak g})=\ch_q {\cal S}({\frak n}^-\oplus{\frak n}^+)
\ch_q {\cal S}({\frak h})=
\ch_q {\cal S}({\frak n}^-\oplus{\frak n}^+)
\ch_q {{\cal S}({\frak g})}^{{\frak g}}
\!\displaystyle\sum_{w\in W}q^{l(w)}.$$
Hence
\begin{equation}\label{hch}
\ch_q H=\ch_q {\cal S}({\frak n}^-\oplus{\frak n}^+)\!\displaystyle
\sum_{w\in W}q^{l(w)}=\prod_{\alpha\in\Delta_0}(1-qe^{\alpha})^{-1}
\ch_q\Lambda {\frak g}_1\!\!\sum_{w\in W}q^{l(w)}.
\end{equation}
\subsection{Classical Case}
\label{hsscl}
We use the previous notations for the corresponding classical objects.
For the classical case the Hesselink formula (see~\cite{jfr}, 9.10) states that
$P_{\lambda}(q)$ is equal to the coefficient of $e^0=1$ in the
expression
\begin{equation}
\label{pseri}
Q_{\lambda}(q)\!:=e^{-\rho}J(e^{\lambda+\rho})\prod_{\alpha\in\Delta^+}
(1-qe^{-\alpha})^{-1}.
\end{equation}
The proof (see~\cite{jfr}, 8.6) is essentially based on the equality
\begin{equation}
\label{nicefrm}
J(e^{\rho}\!\prod_{\alpha\in\Delta^+}(1-qe^{-\alpha}))=J(e^{\rho})\!\!
\sum_{w\in W} q^{l(w)}.
\end{equation}
\subsection{}
A key point of the proof of~\Thm{thmhess} is the following
analogue of equality~\ref{nicefrm}.
\begin{prop}{hesprop}
\begin{equation}
\label{hesfr}
J\left(e^{\rho}\!\biggl(\prod_{\alpha\in{\Delta^+_0}}(1-qe^{-\alpha})\biggr)\!
(1+q^{2l}e^{\beta_1})
\!\!\prod_{\beta\in {\Delta^+_1\setminus\{\beta_1\}}}\!\!(1+e^{\beta})\right)
=J(e^{\rho})\ch_q\Lambda {\frak g}_1\!\!\sum_{w\in W}q^{l(w)}
\end{equation}
where $\beta_1$ is the maximal odd root.
\end{prop}
\begin{pf}
The proposition will be proven in~\ref{heslem}---\ref{endpfprop}.
First of all we need the following technical lemma.
\subsubsection{}
\begin{lem}{heslem}
Set $\Gamma_{\bullet}:=\{\displaystyle\sum_{i=2}^l k_i\beta_i,\ \
k_i\in\{0,1\}\}$.
For any $r=0,1,\ldots, 2l-1$ there exists a unique
$\gamma_r\in \Gamma_{\bullet}$ such that
$\Stab_W(\rho-r\beta_1+\gamma_r)=\{\id\}$. Moreover
$\rho-r\beta_1+\gamma_r=w_r\rho$ for some $w_r\in W$ and $\sgn (w_r)=(-1)^r$.
\end{lem}
\begin{pf}
Recall that the set $\{\beta_i\}_{i=1}^l$ is an orthonormal basis of
${\frak h}^*$ and $W$ acts on this basis by signed permutations.
Therefore the stabilizer of $\lambda\in {\Bbb Q}\pi$ is trivial
iff $|(\lambda,\beta_i)|$ are pairwise distinct non-zero values.
Assume that $\Stab_W(\rho-r\beta_1+\gamma)=\{\id\}$
for some $\gamma\in \Gamma_{\bullet}$ and set $\lambda:=\rho-r\beta_1+\gamma$.
Recall that $(\rho,\beta_i)=l+1/2-i$. Since $0\leq r<2l$ one has
$$\{|(\lambda,\beta_i)|\}_{i=1}^l\subseteq \{l+1/2-i\}_{i=1}^l
=\{|(\rho,\beta_i)|\}_{i=1}^l.$$
Taking into account that all $|(\lambda,\beta_i)|$ are pairwise distinct
we conclude that
$\{|(\lambda,\beta_i)|\}_{i=1}^l=\{|(\rho,\beta_i)|\}_{i=1}^l$.
This implies that $\lambda=w\rho$ for some $w\in W$ and that $\gamma$
is determined by the value of $(\lambda,\beta_1)=l-1/2-r$.
The explicit expressions of $\gamma$ and $w$ for a fixed $r$ are:
$$
\left\{
\begin{array}{lll}
r=0\ & \gamma=0, & w=id \\
1\leq r<l\ & \gamma=\displaystyle\sum_{i=2}^{r+1}\beta_i, & w=s_1\ldots s_r \\
l\leq r<2l\ & \gamma=\displaystyle\sum_{i=2}^{2l-r}\beta_i, & w=s_1\ldots s_l
s_{l-1}\ldots s_{2l-r}
\end{array}
\right.
$$
This completes the proof of the lemma.
\end{pf}
\subsubsection{}
Let $\Delta_{\bullet}$ be the subsystem of $\Delta_0$ generated by
$\pi\setminus\{\beta_1-\beta_2\}$ and $W_{\bullet}$ be the corresponding Weyl
group that is the subgroup of $W$ generated by the reflections
$s_{\alpha_2},\ldots,s_{\alpha_l}$.
Set $\Delta_{\bullet}^+:=\Delta_{\bullet}\cap \Delta^+$,
$\rho_{\bullet}:=\displaystyle\sum_{\alpha\in \Delta_{\bullet}^+}\alpha/2.$
Observe that $\Delta_{\bullet}$ is the root system corresponding
to the simple Lie algebra $\ssp (l-1)$ and that
$\rho_{\bullet}=\rho_0-l\beta_1$. By~\cite{h}, 3.15 for
the Weyl group $W_n$ of the Lie algebra $\ssp(n)$ one has
$\displaystyle\sum_{w\in W_n}q^{l(w)}=
(1-q)^{-n}\displaystyle\prod_{i=1}^n (1-q^{2i})$
and thus
\begin{equation}
\label{qW}
(1-q^{2l})\!\!\sum_{w\in W_{\bullet}} q^{l(w)}=(1-q)\!\!\sum_{w\in W}q^{l(w)}
\end{equation}
\subsubsection{}
\label{endpfprop}
Observe that
\begin{equation}
\label{ex}
\begin{array}{lcl}
(1-q)\displaystyle\prod_{\alpha\in {\Delta_{0}^+}\setminus{\Delta_{\bullet}^+}}
(1-qe^{-\alpha})&=&
(1-qe^{-2\beta_1})(1-q)\displaystyle\prod_{i=2}^l
(1-qe^{-\beta_1-\beta_i})(1-qe^{-\beta_1+\beta_i})\\
&=&\displaystyle\prod_{i=1}^l
(1-qe^{-\beta_1-\beta_i})(1-qe^{-\beta_1+\beta_i})\\
&=&\displaystyle\sum_{r=0}^{2l}\ch (\Lambda^r {\frak g}_1)\!\cdot\!
(-q)^re^{-r\beta_1}.
\end{array}
\end{equation}
The characters $\ch (\Lambda^r {\frak g}_1)$ are stable under
the action of $W$ since ${\frak g}_1$ is a ${\frak g}_0$-module.
By the property (iv) of~\ref{jprop} with $K=W$, one has
\begin{equation}
\label{cc0}
\begin{array}{lcr}
\lefteqn{
(1-q)J\left(e^{\rho}\!\biggl(\displaystyle\prod_{\alpha\in{\Delta^+_0}}
(1-qe^{-\alpha})\biggr)\!
(1+q^{2l}e^{\beta_1})
\!\!\displaystyle\prod_{\beta\in {\Delta^+_1\setminus\{\beta_1\}}}
\!(1+e^{\beta})\right)=}\\
& &\displaystyle\sum_{r=0}^{2l}\ch (\Lambda^r {\frak g}_1)
(-q)^rJ\biggl(e^{\rho-r\beta_1}(1+q^{2l}e^{\beta_1})
\prod_{\alpha\in {\Delta_{\bullet}^+}}(1-qe^{-\alpha})
\prod_{i=2}^l (1+e^{\beta_i})\biggr).
\end{array}
\end{equation}
Decompose $J=J^{W/W_{\bullet}}J^{W_{\bullet}}$ (see (iii) of~\ref{jprop}) and
observe that the elements
$$e^{\beta_1} \text { and }
e^{\rho-\rho_{\bullet}}\displaystyle\prod_{i=2}^l (1+e^{\beta_i})=
e^{l\beta_1}\displaystyle\prod_{i=2}^l (e^{-\beta_i/2}+e^{\beta_i/2})$$
are stable under the action of $W_{\bullet}$.
Using the formula~(\ref{nicefrm}) with respect to $W_{\bullet}$ one can
rewrite the right hand side of~(\ref{cc0}) as
\begin{equation}
\label{cc}
\begin{array}{l}
\displaystyle\sum_{r=0}^{2l}\ch (\Lambda^r {\frak g}_1)
(-q)^rJ^{W/W_{\bullet}}\biggl(e^{\rho-\rho_{\bullet}-r\beta_1}(1+q^{2l}
e^{\beta_1})
\prod_{i=2}^l (1+e^{\beta_i}) J^{W_{\bullet}}\Bigl(e^{\rho_{\bullet}}
\prod_{\alpha\in{\Delta_{\bullet}^+}}(1-qe^{-\alpha})\Bigr)\biggr)=\\
\displaystyle\sum_{r=0}^{2l}\ch (\Lambda^r {\frak g}_1)
(-q)^rJ^{W/W_{\bullet}} \biggl(e^{\rho-\rho_{\bullet}-r\beta_1}(1+q^{2l}
e^{\beta_1})
\prod_{i=2}^l (1+e^{\beta_i})J^{W_{\bullet}}(e^{\rho_{\bullet}})\!\!
\displaystyle\sum_{w\in W_{\bullet}} q^{l(w)}\biggr)=\\
\Bigl(\displaystyle\sum_{w\in W_{\bullet}} q^{l(w)}\Bigr)
\displaystyle\sum_{r=0}^{2l}\ch (\Lambda^r {\frak g}_1)(-q)^r
J\biggl(e^{\rho-r\beta_1}(1+q^{2l}e^{\beta_1})\prod_{i=2}^l
(1+e^{\beta_i})\biggr)=\\
\Bigl(\displaystyle\sum_{w\in W_{\bullet}} q^{l(w)}\Bigr)
\displaystyle\sum_{r=0}^{2l}\ch (\Lambda^r {\frak g}_1)a_r,
\end{array}
\end{equation}
where $a_r\:=(-q)^r
J\biggl(e^{\rho-r\beta_1}(1+q^{2l}e^{\beta_1})\prod_{i=2}^l (1+e^{\beta_i})
\biggr)$.
One can simplify the expression obtained in the following way.
For $r\in\{1,\ldots,2l-1\}$ take $w_r,w_{r-1}$ as in~\Lem{heslem}. Then
\begin{equation}
\label{coeffr}
a_r=(-q)^rJ(e^{w_r\rho}+q^{2l}e^{w_{r-1}\rho})=
(1-q^{2l})q^rJ(e^{\rho}).
\end{equation}
For $r=0,2l$ we have $\ch (\Lambda^0 {\frak g}_1)=
\ch (\Lambda^{2l} {\frak g}_1)=1$,
therefore
$$a_0+a_{2l}=J\left((e^{\rho}+q^{2l}e^{\rho-2l\beta_1})(1+q^{2l}e^{\beta_1})
\displaystyle\prod_{i=2}^l (1+e^{\beta_i})\right).$$
Consider the reflection $s=s_{\beta_1}\in W$.
One has $s\rho=\rho-(2l-1)\beta_1$. Thus the expression
$(e^{\rho+\beta_1}+e^{\rho-2l\beta_1})\displaystyle\prod_{i=2}^l
(1+e^{\beta_i})$ is stable under the action of $s$.
Since $\sgn (s)=-1$, property (i) of~\ref{jprop} implies that
\begin{equation}
\label{cc2}
J\left((e^{\rho+\beta_1}+e^{\rho-2l\beta_1})\prod_{i=2}^l
(1+e^{\beta_i})\right)=0.
\end{equation}
Using~\Lem{heslem} with respect to $r=0$ and $r=2l-1$, we obtain
\begin{equation}
\label{cc1}
J\left((e^{\rho}+q^{4l}e^{\rho-(2l-1)\beta_1})\displaystyle
\prod_{i=2}^l (1+e^{\beta_i})\right)=
J(e^{\rho}-q^{4l}e^{\rho})=J(e^{\rho})(1-q^{2l})(1+q^{2l}).
\end{equation}
Summarizing~(\ref{cc2}) and~(\ref{cc1}) we get
\begin{equation}
\label{coeff0}
a_0+a_{2l}=J\left((e^{\rho}+q^{2l}e^{\rho-2l\beta_1})(1+q^{2l}e^{\beta_1})
\displaystyle\prod_{i=2}^l (1+e^{\beta_i})\right)=J(e^{\rho})(1-q^{2l})
(1+q^{2l}).
\end{equation}
By the substitution of~(\ref{coeffr}) and~(\ref{coeff0}) into~(\ref{cc})
one can rewrite~(\ref{cc0}) as
\begin{equation}
\label{c}
\begin{array}{l}
(1-q)J\left(e^{\rho}\!\biggl(\displaystyle\prod_{\alpha\in{\Delta^+_0}}
(1-qe^{-\alpha})\biggr)\!(1+q^{2l}e^{\beta_1})
\!\!\displaystyle\prod_{\beta\in {\Delta^+_1\setminus\{\beta_1\}}}\!\!
(1+e^{\beta})\right)=\\
\displaystyle\sum_{r=0}^{2l}\ch (\Lambda^r {\frak g}_1)q^rJ(e^{\rho})(1-q^{2l})
\left(\displaystyle\sum_{w\in W_{\bullet}} q^{l(w)}\right)=\\
\displaystyle\sum_{r=0}^{2l}\ch (\Lambda^r {\frak g}_1)q^rJ(e^{\rho})(1-q)
\left(\displaystyle\sum_{w\in W} q^{l(w)}\right)\ \text { by~(\ref{qW})}
\end{array}
\end{equation}
This completes the proof of~\Prop{hesprop}.
\end{pf}
The following theorem is the analogue of Hesselink formula for $\osp(1,2l)$.
\subsection{}
\label{hss2}
\begin{thm}{thmhess}
For any $\lambda\in P^+(\pi)$, the Poincar\'e series
$P_{\lambda}(q)$ is equal to the coefficient of $e^0=1$ in the
expression
\begin{equation}
\label{eqthmhess}
Q_{\lambda}(q)\!:=e^{-\rho}J(e^{\lambda+\rho})\!
\biggl(\prod_{\alpha\in \Delta^+_0}(1-q e^{-\alpha})^{-1}\biggr)\!
(1+q^{2l}e^{-\beta_1})\!\!\prod_{\beta\in
{\Delta^+_1\setminus\{\beta_1\}}}\!\!
(1+e^{-\beta})
\end{equation}
where $\beta_1$ is the maximal odd root.
\end{thm}
\begin{pf}
Observe that $\{\ch {\widetilde V(\lambda)}\!:\lambda\in P^+(\pi)\}$ are
linearly
independent. Hence in order to find the Poincar\'e series $P_{\lambda}(q)$
one can decompose the graded character $\ch_q H$ in terms of
$\{\ch {\widetilde V(\lambda)}\!:\lambda\in P^+(\pi)\}$.
Recall the character formula of
$\widetilde{V}(\lambda),\ \lambda\in P^+(\pi),$ computed by
Kac in~\cite{k1}, (3):
\begin{equation}
\label{kacch}
\ch {\widetilde V}(\lambda)=e^{-\rho}J(e^{\lambda+\rho})
\!\prod_{\alpha\in {\Delta_0^+}}(1-e^{-\alpha})^{-1}
\!\prod_{\beta\in {\Delta_1^+}}(1+e^{-\beta}).
\end{equation}
Therefore
$$J(e^{\rho})\ch_q H=J(e^{\rho})\!\displaystyle\sum_{\lambda\in P^+(\pi)}\!
P_{\lambda}(q)\ch {\widetilde V}(\lambda)=\!\sum_{\lambda\in P^+(\pi)
}P_{\lambda}(q)J(e^{\lambda+\rho}), \text { so }$$
\begin{equation}\label{piq}
P_{\lambda}(q)=\pi_{\lambda+\rho}\bigl(J(e^{\rho})\ch_q H\bigr).
\end{equation}
On the other hand,
\begin{equation}
\begin{array}{lcl}
\label{chqH}
J(e^{\rho})\ch_q H&=&J(e^{\rho})\prod_{\alpha\in\Delta_0}(1-qe^{\alpha})^{-1}
\ch_q\Lambda {\frak g}_1\sum_{w\in W}q^{l(w)} \text { by~(\ref{hch}) }\\
&=&\displaystyle\prod_{\alpha\in\Delta_0}(1-qe^{\alpha})^{-1}
J\left(e^{\rho}\!\biggl(\displaystyle\prod_{\alpha\in
{\Delta^+_0}}(1-qe^{-\alpha})\biggr)
(1+q^{2l}e^{\beta_1})\displaystyle\prod_{i=2}^l (1+e^{\beta_i})\right)\\
&& \hskip 4truecm\text{by ~\Prop{hesprop}}\\
&=&J\left(e^{\rho}\!\biggl(\displaystyle
\prod_{\alpha\in {\Delta^+_0}}(1-qe^{\alpha})^{-1}
\biggr)(1+q^{2l}e^{\beta_1})\displaystyle\prod_{i=2}^l (1+e^{\beta_i})\right)
\\
&& \hskip 4truecm\text{by property (iv) of ~\ref{jprop} with $K=W$.}
\end{array}
\end{equation}
Define the ${\Bbb Z}[[q]]$-automorphism $\iota$
of the ring ${\Bbb Z}[{\frak h}^*][[q]]$ by
$$\iota: e^{\lambda}\mapsto e^{-\lambda},\ \forall \lambda\in{\frak h}^*.$$
Clearly $\pi_0\circ\iota=\pi_0$.
For any $a\in {\Bbb Z}[{\frak h}^*][[q]]$ one has
\begin{equation}
\label{iota}
\begin{array}{lcl}
\pi_{\mu}J(a)&=&\!\displaystyle\sum_{w\in W} \sgn (w)\pi_{w\mu}(a)=\!
\displaystyle\sum_{w\in W} \sgn (w)\pi_0 (e^{-w\mu}a)=
\displaystyle\sum_{w\in W} \sgn (w)\pi_0\circ\iota (e^{-w\mu}a)\\
&=&\pi_0\left(\displaystyle\sum_{w\in W} \sgn (w)(e^{w\mu}\iota(a))\right)=
\pi_0\left(J(e^{\mu})\iota(a)\right).
\end{array}
\end{equation}
It was already shown (see~(\ref{chqH})) that
$$P_{\lambda}(q)=\pi_{\lambda+\rho} J\biggl(e^{\rho}
\displaystyle\prod_{\alpha\in {\Delta^+_0}}(1-qe^{\alpha})^{-1}
(1+q^{2l}e^{\beta_1})\displaystyle\prod_{i=2}^l (1+e^{\beta_i})\biggr).$$
Using~(\ref{iota}) we get
$$P_{\lambda}(q)=\pi_0 \biggl(J(e^{\lambda+\rho})e^{-\rho}
\displaystyle\prod_{\alpha\in {\Delta^+_0}}(1-qe^{-\alpha})^{-1}
(1+q^{2l}e^{-\beta_1})\displaystyle\prod_{i=2}^l (1+e^{-\beta_i})\biggr).$$
This establishes the theorem.\end{pf}
\subsubsection{}
\begin{rem}{remq1}
Observe that $Q_{\lambda}(1)=\ch {\widetilde V}(\lambda)$
(see~(\ref{kacch})). Therefore $P_{\lambda}(1)$ is
equal to the coefficient of $e^0=1$ in $\ch {\widetilde V}(\lambda)$ that
is $\dim {\widetilde V}(\lambda)_0$.
On the other hand, one has
$$P_{\lambda}(1)=\sum_{n=0}^{\infty} [H^n:{\widetilde V}(\lambda)]=
[H:{\widetilde V}(\lambda)]=[{\cal H}:{\widetilde V}(\lambda)]$$
that again gives the second assertion
of the separation theorem
$[{\cal H}:{\widetilde V}(\lambda)]=\dim {\widetilde V}(\lambda)_0$.
\end{rem}
\subsubsection{}
\begin{rem}{hesrem}
One has
$$\pi_{\lambda+\rho}J(a)=\!\displaystyle\sum_{w\in W} \sgn (w)
\pi_{w(\lambda+\rho)}(a)=\!\displaystyle\sum_{w\in W} \sgn (w)
\pi_{w.\lambda}(e^{-\rho}a)$$
where $w.\lambda=w(\lambda+\rho)-\rho$ is a twisted action
of $W$ on ${\frak h}^*$. Thus one can reformulate~\Thm{thmhess}
as follows: for all $\lambda\in P^+(\pi),\ n\in {\Bbb N}$
$$[H^n:\widetilde {V}(\lambda)]=\sum_{w\in W} (-1)^{l(w)} P_n(w.\lambda),$$
where coefficients $P_n(\mu)$ are zero for
$\mu\in P(\pi)\!\setminus\!{\Bbb N}\pi$
and for $\mu\in {\Bbb N}\pi$ are given by the formula
$$
\biggl(\prod_{\alpha\in \Delta^+_0}(1-q e^{\alpha})^{-1}\biggr)
(1+q^{2l}e^{\beta_1})\prod_{\beta\in {\Delta^+_1\setminus\{\beta_1\}}}
(1+e^{\beta})=
\sum_{r=0}^{\infty}\sum_{\nu\in {\Bbb N}\pi}P_r(\nu) e^{\nu} q^r.
$$
\end{rem}
\section {The Parthasarathy--Ranga-Rao--Varadarajan determinants}
\label{prv}
The main result of this Section--~\Thm{thmprv}-- calculates the factorization
of the Parthasarathy--Ranga-Rao--Varadarajan (PRV) determinants.
In~\ref{prvlem},~\ref{prvcons}
we give a construction of these determinants, in~\ref{l1} we study in
details the case $l=1$, and
in~\ref{estimexotic} we compute a bound for the total degrees of the
PRV determinants. The factorization formula for the PRV determinants
is given in~\ref{thmprv}. Subsections~\ref{stfactor}--~\ref{endpfprv}
are devoted to the proof of this formula.
\subsection{}
\label{prvlem}
In order to introduce the Parthasarathy--Ranga-Rao--Varadarajan (PRV)
determinants, start with a little lemma.
Retain notation of~\ref{contrform}. Observe that
$\Upsilon ({\cal U}({\frak g})_{\mu})=0$ for all $\mu\not=0$. In particular,
$\Upsilon(M)=\Upsilon(M_0)$ for any $\ad\frak g$-submodule $M$
of ${\cal U}({\frak g})$.
\subsubsection{}
\begin{lem}{projann}
Let ${M}$ be an $\ad\frak g$-submodule of ${\cal U}({\frak g})$, and
$\widetilde
{V}(\lambda)$ a simple $\frak g$-module of highest weight
$\lambda$, then ${M}\subset \Ann_{{\cal U}({\frak g})}
\widetilde{V}(\lambda)\Longleftrightarrow
\Upsilon({M_0})(\lambda)=0$.
\end{lem}
\begin{pf}
The proof is the same as that of Lemma 7.2 in~\cite{jfr}.
However, we present it here since this is a very important (and quite easy)
fact.
Let $v_{\lambda}$ be a highest weight vector of
$\widetilde{V}(\lambda)$ and
${\widetilde{V}(\lambda)}_-\!:={\frak n}^-{\widetilde{V}(\lambda)}$. One has
$$\Upsilon(M)(\lambda)=0\Longleftrightarrow Mv_{\lambda}\subset
{\widetilde{V}(\lambda)}_- .$$
On the other hand, from the $\ad{\frak g}$-invariance of $M$ one has
${\cal U}({\frak n}^-)M=M{\cal U}({\frak n}^-)$ and so the assumption
$\Upsilon(M_0)(\lambda)=0$ implies
$$M\widetilde{V}(\lambda)=M{\cal U}({\frak n}^-)v_{\lambda}
={\cal U}({\frak n}^-)Mv_{\lambda}\subset {\cal U}({\frak n}^-)
{\widetilde{V}(\lambda)}_-\subset
{\widetilde{V}(\lambda)}_-\ssubset{\widetilde{V}(\lambda)}.$$
The $\ad{\frak g}$-invariance of $M$ implies that
${\cal U}({\frak g})M=M{\cal U}({\frak g})$ and consequently
$M\widetilde{V}(\lambda)$ is a submodule of $\widetilde{V}(\lambda)$.
Hence $M{\widetilde{V}(\lambda)}=0$.
\end{pf}
\subsection{The construction of the PRV determinants}
\label{prvcons}
The separation theorem leads to the following contruction.
Fix a simple finite dimensional module
$\widetilde{V}(\lambda)$ of the highest weight $\lambda\in P^+(\pi)$
and let $C(\lambda)$ be the corresponding isotypical component in ${\cal H}$.
\begin{equation}
\label{C}
C(\lambda)=\displaystyle\mathop{\oplus}_{j=1}^{\dim {V(\lambda)_0}} V^{(j)},
\ \ \widetilde{V}(\lambda)\iso V^{(j)}.
\end{equation}
Choose a basis $\{v_i\}$ of $\widetilde{V}(\lambda)_0$ and let $\{v_i^j\}$
be the corresponding basis of $V^{(j)}_0$. Consider the matrix
$PRV^{\lambda}:=(\Upsilon(v_i^j))$.
\Lem{projann} implies that the $j^{th}$ column of the matrix
$PRV^{\lambda}$ is zero at a point $\mu\in\fh^*$
iff $V^{(j)}\subset \Ann \widetilde{V}(\mu)$. Notice that the
matrix $PRV^{\lambda}$ depends on the choice of the decomposition
of the isotypical component $C(\lambda)$, on the choice of the
basis of $\widetilde{V}(\lambda)_0$
and on the choice of the isomorphisms from each $V^{(j)}$ to
$\widetilde{V}(\lambda)$.
For any change of these parameters, the new matrix one obtains
is of the form $M(PRV^{\lambda})N$ where $M,N$ are invertible square
complex matrices. Therefore, the corank of the
$PRV^{\lambda}$ is correctly defined and
$\det PRV^{\lambda}\in {\cal S}({\frak h})$
is defined up to a non-zero scalar. Hence,
for any $\mu\in {\frak h}^*$, one has the equivalence:
$$\det {PRV}^{\lambda}(\mu)=0\Longleftrightarrow
C(\lambda)\cap \Ann_{\cal H}\widetilde{V}(\mu)\not=0.$$
Moreover
\subsubsection{}
\begin{lem}{clprv}
Take $\lambda\in P^+(\pi)$ and $\mu\in {\frak h}^*$. Then
\begin{enumerate}
\item $\corank PRV^{\lambda}(\mu)=[\Ann_{\cal H}\widetilde{V}(\mu):\widetilde
{V}(\lambda)]$
\item $\mu$ is a zero of
$\det PRV^{\lambda}$ of order $\geq [\Ann_{\cal H}\widetilde{V}(\mu):
\widetilde{V}(\lambda)]$
\item for any $\lambda\in P^+(\pi)$,
the polynomial $\det PRV^{\lambda}$ is not identically zero.
\end{enumerate}
\end{lem}
\begin{pf} From the comments above (i) and (ii) are straightforward.
The proof of (iii) coincides with the proof in 9.11.4,~\cite{jfr}.
\end{pf}
\subsection{Case $l=1$}
\label{l1}
This case was completely described by Pinczon in~\cite{pin}.
In this case $\pi=\Delta_1^+=\{\beta\},\
\Delta_0^+=\{2\beta\},\ \rho={1\over 2}\beta,\ {\frak h}^*={\Bbb C}\beta,
{\frak h}={\Bbb C}\varphi(\beta),\ P^+(\pi)={\Bbb N}\beta$.
\subsubsection{}
\label{odd}
Denote by ${\widetilde V}(n)$ a simple module of the highest weight $n\beta$.
Call a simple module $V$ {\em odd} if $V\cong {\widetilde V}(n)$
for some odd number $n$.
Since $\dim {\widetilde V}(m)_0=1$ for all $m\in {\Bbb N}$ the separation
theorem implies that
$${\cal H}={\underset{m=0}{\overset{\infty}\oplus}}{\widetilde V}(m).$$
Taking into account that for any $m\in {\Bbb N},\
s_{\beta}.(m\beta)\!=\!-(m+1)\beta\not\in P^+(\pi)$,
we conclude from~\Rem{hesrem} that
$$[H^n:\widetilde {V}(m)]=P_n(m\beta)$$
where the coefficients $P_n(m\beta)$ are given by the formula
$$\displaystyle\sum_{n=0}^{\infty}\displaystyle\sum_{m=0}^{\infty}P_n(m\beta)
e^{m\beta}q^n=
(1-qe^{2\beta})^{-1}(1+q^2e^{\beta})=1+qe^{2\beta}+
\displaystyle\sum_{n=2}^{\infty}q^n(e^{2n\beta}+e^{(2n-3)\beta}).$$
Hence
\begin{equation}
H^n=\widetilde {V}(2n)\oplus\widetilde {V}(2n-3)\mbox{ for }n\geq 2
\end{equation}
Since ${\cal S}({\frak g})={ H}\otimes {{\cal S}({\frak g})}^{\frak g}$,
we conclude that for $m=2n, 2n-3$ one has
$[{\cal S}^k({\frak g}):\widetilde {V}(m)]=0$ for $0\leq k<n$
and $[{\cal S}^n({\frak g}):\widetilde {V}(m)]=1$. Hence
${\cal F}^n$ contains a unique copy of $\widetilde {V}(2n)$ and
$\widetilde {V}(2n-3)$ which lie in ${\cal H}$.
This shows in particular that, for any $n\geq 0$, the degrees of
$\det PRV^{2n\beta}$ and $\det PRV^{2n+1}$ are less than or equal to
$n$ and $n+2$ respectively.
\subsubsection{}
\label{prvl1}
Let $V$ be the copy of $\widetilde{V}(2n)$ in ${\cal H}$ and $v$ its
highest weight vector. For any $k\geq 0$,
$V\subset \Ann_{\cal H} \widetilde{V}(k)$ iff $v\widetilde{V}(k)=0$. As
$\Omega({\widetilde{V}(k)})=\{k,k-1,\ldots,0,\ldots,-k+1,-k\}$,
the condition $n>k$ is
sufficient to ensure that $v\widetilde{V}(k)=0$ and therefore that
$V\subset \Ann_{\cal H} \widetilde{V}(k)$.
So $\det PRV^{2n\beta}$ is divisible by
$\mathop{\prod}_{k=0}^{n-1}(\varphi(\beta)-k)$.
Analogously we obtain also the divisibility of $\det PRV^{(2n+1)\beta}$ by
$\mathop{\prod}_{k=0}^{n}(\varphi(\beta)-k)$.
Taking into account that the degree of $\det PRV^{2n\beta}\leq n$
we conclude that, up to a non-zero scalar,
$$\det PRV^{2n\beta}=\mathop{\prod}_{k=0}^{n-1}(\varphi(\beta)-k).$$
\subsubsection{}
The factorization of $\det PRV^{(2n+1)\beta}$ requires some extra work.
The Casimir operator $C_0$ of ${\cal U}({\frak g}_0)$
acts on a ${\frak g}_0$-Verma module $M(\mu)$ of highest weight $\mu$
by the scalar $(\mu,\mu+2\rho_0)$.
Since ${\widetilde M}(\mu)=M(\mu)\oplus M(\mu-\beta)$ as a
${\frak g}_0$-module,
it follows that $C_0$ acts on ${\widetilde M}(\mu)$ by a scalar
iff $\mu=-\beta/2=-\rho$.
Hence $(C_0+{3\over 4})\in\Ann {\widetilde M}(-\rho)$.
Therefore $\Ann {\widetilde M}(-\rho)$ contains $V\!:=\ad {\cal U}({\frak g})
(C_0+{3\over 4})$.
Since ${\frak g}_0$ acts trivially on $(C_0+{3\over 4})$ and
$\Omega({\frak g}_1)=\{\beta,-\beta\}$, it follows that
$\ V\subseteq V^{(1)}\oplus V^{(2)}$ where $V^{(1)}\cong{\widetilde V}(1),\
V^{(2)}\cong {\widetilde V}(0).$
On the other hand, $C_0\not\in {\cal Z}({\frak g})$ so
$V^{(2)}\not=V$. Hence $V^{(1)}\subseteq V$.
Since $C_0\in {\cal F}^2$, we conclude
that $\Ann {\widetilde M}(-\rho)$ contains a copy $V^{(1)}$ of
${\widetilde V}(1)$
and this copy lies in ${\cal F}^2$. Thus $V^{(1)}$ lies in ${\cal H}$.
Fix a primitive vector $a$ of $V^{(1)}$. Taking into account that
$X_{2\beta}$ is a primitive vector of ${\frak g}\cong{\widetilde V}(2)$
and ${\cal U}({\frak g})$ is a domain, we conclude that the product
$X_{2\beta}^ka$ is also a primitive vector of the weight $(2k+1)\beta$ and so
$\ad {\cal U}({\frak g})(X_{2\beta}^ka)\cong{\widetilde V}(2k+1)$.
Since $X_{2\beta}^ka\in {\cal F}^{k+2}$ and $a\in \Ann {\widetilde M}(-\rho)$,
it follows that $\ \ad {\cal U}({\frak g})(X_{2\beta}^ka)\subset
\Ann_{\cal H} {\widetilde M}(-\rho)$. Hence
\begin{equation}
\label{annl1}
\Ann_{\cal H} {\widetilde M}(-\rho)\supseteq
{\underset{k=0}{\overset{\infty}\oplus}}
{\widetilde V}(2k+1).
\end{equation}
Since $W.(-\rho)=\{-\rho\}$, the module ${\widetilde M}(-\rho)$ is simple.
So (\ref{annl1}) means that $\varphi(\beta)+{1\over 2}$ divides all
$\det PRV^{(2n+1)\beta}$, $n\geq 0$.
We conclude from~\ref{prvl1} that, up to a non-zero scalar,
$$\det PRV^{(2n+1)\beta}=(\varphi(\beta)+{1\over 2})
\mathop{\prod}_{k=0}^{n}(\varphi(\beta)-k).$$
From the factorization of $\det PRV^{2n\beta}$ and $\det PRV^{(2n+1)\beta}$
we derive that
(\ref{annl1}) is an equality and that $\Ann_{\cal H} {\widetilde M}(\mu)=0$
if $\widetilde{M}(\mu)$ simple and $\mu\not=-\rho$.
Since $W.(-\rho)=\{-\rho\}$,
$\Hom_{\frak g}(\widetilde{M}(-\rho),\widetilde{M}(\mu))=0$ for any
$\mu\not=-\rho$.
Hence $\Ann_{\cal H} {\widetilde M}(\mu)=0$ for any $\mu\not=-\rho$.
We summarize:
\begin{eqnarray*} \mu\not=-\rho & \Longrightarrow &
\Ann_{\cal H}\widetilde M(\mu)=0\\
\mu=-\rho & \Longrightarrow & \Ann_{\cal H}\widetilde M(-\rho)
={\underset{i=0}{\overset{\infty}\oplus}}
{\widetilde V}(2i+1)
\end{eqnarray*}
In the sequel we will need the following
\subsubsection{}
\begin{lem}{lemcasel1}
Take ${\frak g}=\osp(1,2)$. For any odd submodule $V$ of
${\cal U}({\frak g})$ (see~\ref{odd}) and any $v\in V$, the Harish-Chandra
projection $\Upsilon (v)$ is divisible by $(\varphi(\beta)+{1\over 2})$.
\end{lem}
\begin{pf}
Since ${\cal Z}({\frak g})$ acts on ${\widetilde M}(-\rho)$ by scalars,
the separation theorem implies that as ${\frak g}$-module
$${\cal U}({\frak g})/(\Ann {\widetilde M}(-\rho))\cong{\cal H}/
(\Ann_{\cal H} {\widetilde M}(-\rho))={\underset{i=0}{\overset{\infty}\oplus}}
{\widetilde V}(2i).$$
Thus $\Ann {\widetilde M}(-\rho)$ contains all odd submodules of
${\cal U}({\frak g})$.
By~\Lem{projann}, $\Upsilon (v)(-\rho)=0$ for all $v\in V$.
Since $\Upsilon (v)$ is just a polynomial of one variable, it follows that
$\Upsilon (v)$ is divisible by
$\varphi(\beta)+(\rho,\beta)=\varphi(\beta)+{1\over 2}$ as required.
\end{pf}
\subsection{A bound for the total degrees of PRV determinants}
\label{estimexotic}
Retain notations of~\ref{prvcons}.
Fix a finite dimensional module ${\widetilde V}(\lambda)$ and
a decomposition (\ref{C}) of
its isotypical component $C(\lambda)$ in ${\cal H}$.
For each copy $V^{(i)}\subset {\cal H}$ let $n_i$ be such that
$\gr_{\cal F} V^{(i)}\subset H^{n_i}$.
The subspaces ${\cal F}^{(k)}$ are stable under the projection
$\Upsilon$ so
$\Upsilon(V^{(i)})\subset {\cal F}^{(n_i)}\cap {\cal U}({\frak h})$.
Thus any entry $\Upsilon(v^i_j)$ of $i$-th column of $PRV^{\lambda}$
has degree (as polynomial in $S({\frak h})$) less than or equal to $n_i$.
Using prime to denote the derivative with respect to $q$ we obtain
$$\text {degree }\det PRV^{\lambda}\leq\displaystyle
\sum_{r=0}^{\infty} r[H^r:{\widetilde V}(\lambda)]=P'_{\lambda}(1)$$
where $\det PRV^{\lambda}$ is considered as a polynomial in $S({\frak h})$.
Using notations of~\ref{nthess} and~\Thm{thmhess}
we can rewrite~(\ref{eqthmhess}) as $P_{\lambda}(q)=\pi_0(Q_{\lambda}(q))$.
Hence
\begin{equation}
\label{estimth}
\text {degree }\det PRV^{\lambda}\leq\pi_0(Q_{\lambda}'(1)).
\end{equation}
Taking the derivative of~(\ref{eqthmhess}) on $q$, we obtain
$$Q_{\lambda}'(q)=Q_{\lambda}(q)\left(\displaystyle\sum_{\alpha\in\Delta_0^+}
e^{-\alpha}(1-qe^{-\alpha})^{-1}+2lq^{2l-1}e^{-\beta_1}
(1+q^{2l}e^{-\beta_1})^{-1}
\right).$$
By~\Rem{remq1}, $Q_{\lambda}(1)=\ch {\widetilde V}(\lambda)$. Thus
\begin{equation}
\label{estimprv1}
\begin{array}{lcl}
Q_{\lambda}'(1)&=&Q_{\lambda}(1)\left(\displaystyle\sum_{\alpha\in\Delta_0^+}
e^{-\alpha}(1-e^{-\alpha})^{-1}+2le^{-\beta_1}(1+e^{-\beta_1})^{-1}\right)\\
&=&\displaystyle\sum_{\alpha\in\Delta^+_0}\displaystyle
\sum_{m=1}^{\infty}e^{-m\alpha}\ch {\widetilde V}(\lambda)+
2l\displaystyle\sum_{m=1}^{\infty} (-1)^{m+1}e^{-m\beta_1}
\ch {\widetilde V}(\lambda)
\end{array}
\end{equation}
Since $\pi_0(e^{-\mu}\ch {\widetilde V}(\lambda))
=\dim {\widetilde V}(\lambda)_{\mu}\,$ for any $\mu\in P(\pi)$,
the formulas~(\ref{estimth}),~(\ref{estimprv1}) imply
\begin{equation}
\label{estimprv}
\text {degree }\det PRV^{\lambda}\leq \displaystyle\sum_{m=1}^{\infty}
\left(\displaystyle\sum_{\alpha\in\Delta^+_0}
\dim \widetilde {V}(\lambda)_{m\alpha}+
2l(-1)^{m+1}\dim \widetilde {V}(\lambda)_{m\beta_1}\right)
\end{equation}
\subsection{Factorization of PRV determinants} Let us formulate the main
theorem of this Section:
\begin{thm}{thmprv} For all $\lambda\in P^+(\pi)$, one has up to a non-zero
scalar,
$\vtop{\hbox to \textwidth{$\det
{PRV}^{\lambda}=\underbrace{\displaystyle\mathop{\prod}_
{\alpha\in\overline{\Delta_0^+}}
\mathop{\prod}_{m=1}^{\infty}
{\bigl[\varphi(\alpha)+(\rho,\alpha)-{1\over
2}m(\alpha,\alpha)\bigr]}^{\dim{\widetilde{V}(\lambda)}_{m\alpha}}}_{\mbox
{\footnotesize standard
factors}}\times$\hfill}
\hbox to\textwidth{\hfill$\underbrace{\displaystyle\mathop{\prod}
_{\alpha\in{\Delta_1^+}}
\mathop{\prod}_{m=1}^{\infty}
{\bigl[\varphi(\alpha)+(\rho,\alpha)-{1\over2}(2m-1)(\alpha,\alpha)\bigr]}^
{dim{\widetilde{V}(\lambda)}_{(2m-1)\alpha}}}_{\mbox{\footnotesize
standard factors}}\ \times $\hfill}
\hbox to \textwidth{\hfill$\underbrace{\displaystyle\mathop{\prod}_{\alpha\in
\Delta_1^+}
{\bigl[\varphi(\alpha)+(\rho,\alpha)\bigr]}^{\mathop{\sum}_{i=1}^{\infty}
{(-1)}^{i+1}
\dim {\widetilde{V}(\lambda)}_{i\alpha}}}_
{\mbox{\footnotesize exotic factors}}$}}$
\end{thm}
The rest of this section is devoted to the proof of this theorem.
\subsection{Standard factors of the PRV determinants}
\label{stfactor}
In the subsections~\ref{stfactor}---~\ref{pfthmLL} we shall explain
briefly how
certain factors of the PRV determinants can be obtained
following the procedure used by Joseph in the classical case~\cite{jfr},
Chapter 9. These factors will be called "standard factors".
Let first fix some notations. Let $M=M_0\oplus M_1$ be a ${\frak g}$-module.
We endow the
space $\Hom_{{\Bbb C}}(M,M)$ of ${\Bbb C}$-linear endomorphisms of $M$
with the natural supervector space structure.
The ${\frak g}$-module structure on
$\Hom_{{\Bbb C}}(M,M)$ is defined on homogenous elements $x\in {\frak g},\
f\in \Hom_{{\Bbb C}}(M,M)$ by the formula
$$ (xf)(m)=x(f(m))-{(-1)}^{|x||f|}f(xm)$$
The locally finite part of $\Hom_{{\Bbb C}}(M,M)$ with respect to this
${\frak g}$-module structure
will be denoted $\F(M,M)$. For the above ${\frak g}$-module structure on
$\Hom_{{\Bbb C}}(M,M)$ the natural map
${\cal U}({\frak g})\longrightarrow \Hom_{{\Bbb C}}(M,M)$ is
a ${\frak g}$-module map for the adjoint action on ${\cal U}({\frak g})$.
The image is contained in $\F(M,M)$.
Take $\mu\in {\frak h}^*$ and consider the induced morphism
$${\cal H}\slash \Ann_{\cal H}\widetilde{V}(\mu)\longrightarrow
\F(\widetilde{V}(\mu),\widetilde{V}(\mu)).$$
The idea to find the "standard zeroes" of $\det PRV^{\lambda}$ is to
construct some $\mu$ such that the
multiplicity of $\widetilde{V}(\lambda)$ in
$\F(\widetilde{V}(\mu),\widetilde{V}(\mu))$ is less than the multplicity of
$\widetilde{V}(\lambda)$
in ${\cal H}$.
Define for $\alpha\in \Delta_{irr}$,
$$\Lambda_{m,\alpha}\!:=\{\mu\in {\frak h}^*|\ \langle\mu+\rho,\alpha\rangle
=m,\ \ \langle\mu+\rho,\beta\rangle\not\in {\Bbb Z}\ ,\forall
\beta\in \Delta^+_{irr}\backslash
\{\alpha\}\}$$
for all $m\in {\Bbb N}^+$ if $\alpha\in \overline{\Delta_0^+}$ and for all odd
$m$ if $\alpha\in \Delta_1^+$. For such choices of $m$ and $\alpha$, the set
$\Lambda_{m,\alpha}$ is obviously non-empty.
One has the
\subsubsection{}
\begin{thm}{thmLambda} Let $\lambda\in P^+(\pi)$, $\alpha\in \Delta^+_{irr}$,
$m\in {\Bbb N}^+$ (assumed
to be odd if $\alpha$ is odd), $\mu\in \Lambda_{m,\alpha}$ then
$$\bigl[\F(\widetilde{V}(\mu),\widetilde{V}(\mu)):\widetilde{V}(\lambda)\bigr]=
\dim {\widetilde{V}(\lambda)}_0- \dim {\widetilde{V}(\lambda)}_{m\alpha}.$$
\end{thm}
Assume $m,\alpha$ chosen as above. Then by~\Lem{clprv} and the theorem,
$\mu\in \Lambda_{m,\alpha}$ is a zero of $\det PRV^{\lambda}$ of order
$\geq \dim {\widetilde{V}(\lambda)}_{m\alpha}$. The same holds for
the Zariski closure of
$\Lambda_{m,\alpha}$, namely for the hyperplane $(\mu+\rho,\alpha)-
m\displaystyle {(\alpha,\alpha)\over 2}$.
This means that $\bigl[\varphi(\alpha)+(\rho,\alpha)-m\displaystyle
{(\alpha,\alpha)\over 2}\bigr]$ divides $\det PRV^{\lambda}$.
Consequently we obtain the
\subsubsection{}
\begin{cor}{stfact} For any $\lambda\in P^+(\pi)$, the determinant
$\det PRV^{\lambda}$ is divisible by
$\vtop{\hbox to
\textwidth{$\quad \displaystyle\mathop{\prod}_{\alpha\in\overline{\Delta_0^+}}
\mathop{\prod}_{m=1}^{\infty}
{\bigl[\varphi(\alpha)+(\rho,\alpha)-{1\over
2}m(\alpha,\alpha)\bigr]}^{\dim{\widetilde{V}(\lambda)}_{m\alpha}}\times$
\hfill}
\hbox to \textwidth{\hfill$\displaystyle\mathop{\prod}_{\alpha\in{\Delta_1^+}}
\mathop{\prod}_{m=1}^{\infty}
{\bigl[\varphi(\alpha)+(\rho,\alpha)-{1\over2}(2m-1)(\alpha,\alpha)\bigr]}^
{dim{\widetilde{V}(\lambda)}_{(2m-1)\alpha}}\hfill$}}$
\end{cor}
\subsection{Sketch of the proof of~\Thm{thmLambda}}
\label{pfthmLL} This subsection being very close to the
proposition 9.7,~\cite{jfr}, we present only a
sketch of the proof. First comes a
\subsubsection{}
\begin{defn}{defdom} We call $\lambda\in {\frak h}^*$ dominant if
$$\begin{array}{lll} \forall \alpha\in \overline{\Delta_0^+},&&
\langle\lambda+\rho,\alpha\rangle\not\in -{\Bbb N}^+ \\
\forall \alpha\in \Delta_1^+,&& \langle\lambda+\rho,\alpha\rangle\not\in
-2{\Bbb N}-1\end{array}$$
\end{defn}
One has the key
\subsubsection{}
\begin{lem}{projectivity} The following conditions on $\lambda$ are
equivalent
\begin{enumerate}
\item $\widetilde{M}(\lambda)$ is projective in $\widetilde{\cal O}$
\item $\forall \mu\in{\frak h}^*,\ \Hom_{\frak g}(\widetilde{M}(\lambda),
\widetilde{M}(\mu))\not=0
\Longrightarrow \mu=\lambda$
\item $\lambda$ is dominant
\item $\forall \mu\in{\frak h}^*,\
[\widetilde{M}(\mu):\widetilde{V}(\lambda)]\not=0
\Longrightarrow \mu=\lambda$.
\end{enumerate}
\end{lem}
One deduce from the lemma the following
\subsubsection{}
\begin{prop}{propdom} Let $\mu$ be dominant, $\mu'\in {\frak h}^*$ and $\lambda
\in P^+(\pi)$. Then
$$\bigl[\F(\widetilde{M}(\mu),\widetilde{M}(\mu')):
\widetilde{V}(\lambda)\bigr]=
\dim{\widetilde{V}(\lambda)}_{\mu-\mu'}$$
\end{prop}
\subsubsection{}
Retain notations of~\Thm{thmLambda}. Recall that $\mu\in\Lambda_{m,\alpha}$
and hence is dominant.
According to factorization of Shapovalov determinants given
in~\Thm{shapovalov}, $\widetilde{M}(\mu)$ contains
a unique copy of $\widetilde{M}(s_{\alpha}.\mu)$.
Moreover we have
the short exact sequence
$$ 0\longrightarrow \widetilde{M}(s_{\alpha}.\mu)\longrightarrow
\widetilde{M}(\mu)\longrightarrow
\widetilde{V}(\mu)\longrightarrow 0$$ and $\widetilde{M}(s_{\alpha}.\mu)$
is a simple Verma module. An easy consequence of the
lemma~\ref{projectivity} is the exactness of the functor
$\F(\widetilde{M}(\mu),-)$. This provides
the exact sequence
$$ 0\longrightarrow \F\bigl(\widetilde{M}(\mu),\widetilde{M}(s_{\alpha}.\mu)
\bigr)\longrightarrow
\F\bigl(\widetilde{M}(\mu),\widetilde{M}(\mu)\bigr)\longrightarrow
\F\bigl(\widetilde{M}(\mu),\widetilde{V}(\mu)\bigr)\longrightarrow 0$$
One the other hand, the functor $\F\bigl(-,\widetilde{V}(\mu)\bigr)$ gives
the exact sequence
$$ \F\bigl(\widetilde{M}(s_{\alpha}.\mu),\widetilde{V}(\mu)\bigr)
\longleftarrow
\F\bigl(\widetilde{M}(\mu),\widetilde{V}(\mu)\bigr)\longleftarrow
\F\bigl(\widetilde{V}(\mu),\widetilde{V}(\mu)\bigr)\longleftarrow 0$$
An argument from Gelfand-Kirillov dimension theory shows
that $ \F\bigl(\widetilde{M}(s_{\alpha}.\mu),\widetilde{V}(\mu)\bigr)=0$.
The theorem then results from the proposition~\ref{propdom}.
\subsection{"Exotic" factors of PRV determinant}
Call the factors of the PRV determinants "exotic" if they are not standard.
By~\Cor{stfact}, the total degree of the product of the standard factors
of $\det PRV^{\lambda}$ is at least
$$\sum_{m=1}^{\infty}
\left(\displaystyle\sum_{\alpha\in {\overline\Delta^+}_0}
\dim \widetilde {V}(\lambda)_{m\alpha}+
\displaystyle\sum_{\beta\in {\Delta_1^+}}
\dim \widetilde {V}(\lambda)_{(2m-1)\beta}\right).$$
Recall that ${\Delta_1^+}\subset W\beta_1$ and so
$\dim\widetilde {V}(\lambda)_{k\beta}=\dim\widetilde {V}(\lambda)_{k\beta_1}$
for all $\beta\in {\Delta_1^+}$.
Taking into account that $\Delta^+_0=2\Delta^+_1\cup{\overline\Delta^+}_0$,
we conclude from~(\ref{estimprv}) that
the total degree of the product of the exotic factors
of $\det PRV^{\lambda}$ is less than or equal to
\begin{equation}
\label{estimprvex}
\sum_{\beta\in {\Delta_1^+}}\sum_{m=1}^{\infty}(-1)^{m+1}\dim
\widetilde {V}(\lambda)_{m\beta}.
\end{equation}
\subsubsection{}\label{remprvl1}
\begin{rem}{} Consider the case $l=1$, i.e. ${\frak g}=\osp(1,2)$
and retain notations of~\ref{l1}.
Then the formula (\ref{estimprvex}) takes the values
\begin{equation}
\label{l1prvex}
\sum_{m=1}^{\infty}(-1)^{m+1}\dim {\widetilde V}(n)_{m\beta}=
\left\{
\begin{array}{ll}
0,&\ n \ \text { is even}\\
1,&\ n \ \text { is odd}
\end{array}
\right.
\end{equation}
\end{rem}
\subsection{}
\label{genl}
Let ${\frak p}$ be
$${\frak p}\!:={\frak g}_{\pm\beta_l}\oplus {\frak g}_{\pm2\beta_l}\oplus
{\Bbb C}\varphi(\beta_l).$$
Then ${\frak p}$ is a Lie subsuperalgebra of ${\frak g}$ and
there is an obvious isomorphism ${\frak p}\cong\osp(1,2)$ which
maps the element $\varphi(\beta_l)$ to $\varphi(\beta)$.
We shall denote a simple
${\frak p}$-module of the highest weight $n\beta_l$ by ${\breve V}(n)$ and
call a simple ${\frak p}$-module $V$ "odd" if $V\cong{\breve V}(n)$ for some
odd number $n$.
Lemma~\ref{lemcasel1} has the following useful generalization.
\subsubsection{}
\begin{lem}{lembeta}
For any odd $\ad {\frak p}$-submodule $V$ of ${\cal U}({\frak g})$
and any $v\in V$, the Harish-Chandra
projection $\Upsilon (v)$ is divisible by $\varphi(\beta_l)+{1\over 2}=
(\varphi(\beta_l)+(\rho,\beta_l))$.
\end{lem}
\begin{pf}
Since $\beta_l$ is a simple root of $\Delta$,
$({\frak p}+{\frak b}^{\pm})$ are
Lie subsuperalgebras and even parabolic subsuperalgebras of ${\frak g}$.
Let ${\frak h}'$ be the linear span of $\{\varphi(\beta_i)\}_{i=1}^{l-1}$.
The superalgebra $({\frak p}\oplus{\frak h}')$ is the Levi factor
of the parabolic superalgebras $({\frak p}+{\frak b}^{\pm})$.
Let ${\frak m}^{\pm}$ be the nilradical of $({\frak p}+{\frak b}^{\pm})$.
Denote by ${\cal U}({\frak p}+{\frak h}')$ the universal enveloping
algebra of $({\frak p}+{\frak h}')$.
Let $P$ be the projection of ${\cal U}({\frak g})$ onto
${\cal U}({\frak p}+{\frak h}')$
with respect to decomposition
${\cal U}({\frak g})={\cal U}({\frak p}+{\frak h}')\oplus
({\frak m}^-{\cal U}({\frak g})+{\cal U}({\frak g}){\frak m}^+).$
Then $\Upsilon=\Upsilon\circ P$.
Fix a simple odd $\ad {\frak p}$-module $V\subset {\cal U}({\frak g})$ and
$v\in V$. Since
$P$ is an $\ad {\frak p}$-homomorphism, it follows that either $P(V)=0$ or
$P(V)\simeq V$. In the first case the assertion obviously holds. Consider the
second case. Since
${\frak p}$ acts trivially on ${\frak h'}$ and on its universal enveloping
algebra ${\cal U}({\frak h'})$, the multiplication map induces an isomorphism
${\cal U}({\frak p}+{\frak h}')={\cal U}({\frak p})\otimes
{\cal U}({\frak h'})$ as ${\frak p}$-modules.
Write $P(v)=\sum p_ih_i$ with non-zero $p_i\in {\cal U}({\frak p}),\
h_i\in {\cal U}({\frak h'})$. Since $\ad {\cal U}({\frak p})P(v)\!\simeq\!V$
is a simple ${\frak p}$-module, one can suppose that
$V\!\simeq\!\ad {\cal U}({\frak p})(p_i)$ for all $i$.
Since $p_i\in {\cal U}({\frak p})$, $\Upsilon(p_i)$
is divisible by $(\varphi(\beta_l)+{1\over 2})$ for all $i$ by~\Lem{lemcasel1}.
Thus $\Upsilon(v)=\Upsilon(P(v))=\sum \Upsilon(p_ih_i)=\sum \Upsilon(p_i)h_i$
is divisible by $(\varphi(\beta_l)+{1\over 2})$ as required.
This establishes the lemma.
\end{pf}
\subsubsection{}
\label{casebetan}
Fix $\lambda\in P^+(\pi)$ and an $\ad{\frak g}$-submodule
$V\cong \widetilde {V}(\lambda)$ of ${\cal H}$. Then
there is an $\ad{\frak p}$-submodule of $V$ given by the formula
$${\breve V}\!:={\underset{m\in {\Bbb Z}}\oplus} V_{m\beta_l}.$$
Fix a decomposition of ${\breve V}$ into the sum of simple
$\ad{\frak p}$-modules.
Taking into account that $V_0\subset {\breve V}$ and that
the dimension of the zero weight space of any simple ${\frak p}$-module
is equal to one, we conclude that the number of simple $\ad{\frak p}$-modules
in the decompositon of ${\breve V}$ is equal to $\dim V_0$. Hence
$${\breve V}={\underset{i=1}{\overset{\dim V_0}\oplus}} {\breve V}^{(i)}.$$
Using~\ref{remprvl1} we obtain
\begin{equation}
\label{alsum}
\displaystyle\sum_{m=1}^{\infty}(-1)^{m+1}
\dim {\widetilde{V}(\lambda)}_{m\beta_l}=
\displaystyle\sum_{i=1}^{\dim {V}_0} \displaystyle\sum_{m=1}^{\infty}
(-1)^{m+1}\dim {\breve V}^{(i)}_{m\beta_l}=
\# \{i:\ {\breve V}^{(i)} \text { is odd }\}.
\end{equation}
Choose a basis $\{v_i\}_{i=1}^{\dim V_0}$ of the subspace $V_0$ such
that $v_i\in {\breve V}^{(i)}$ for all $i=1,\ldots,\dim V_0$. Consider
a matrix $PRV^{\lambda}$ constructed using this basis. By~\Lem{lembeta}
the polynomial $\Upsilon (v_i)$ is divisible by
$(\varphi(\beta_l)+(\rho,\beta_l))$ if ${\breve V}^{(i)}$ is an odd
${\frak p}$-module.
Thus all entries of the $i$-th line of the matrix $PRV^{\lambda}$ are
divisible by $(\varphi(\beta_l)+(\rho,\beta_l))$ if ${\breve V}^{(i)}$
is an odd ${\frak p}$-module. Using the equality~(\ref{alsum}) we obtain the
\subsubsection{}\begin{cor}{corbeta}
(i) For any $\lambda\in P^+(\pi)$ and any $\mu\in {\frak h}^*$ such that
$(\mu+\rho,\beta_l)=0$ one has
$$\corank PRV^{\lambda}(\mu)\geq \sum_{m=1}^{\infty}(-1)^{m+1}
\dim {\widetilde V}(\lambda)_{m\beta_l}.$$
(ii) For any $\lambda\in P^+(\pi)$, the determinant $\det PRV^{\lambda}$ is
divisible by
$$\left(\varphi(\beta_l)+(\rho,\beta_l)\right)^{\sum_{m=1}^{\infty}(-1)^{m+1}
\dim {\widetilde V}(\lambda)_{m\beta_l}}$$
\end{cor}
\subsection{Democracy principle}
\label{democrasy}
In~\Cor{corbeta} we found
$m_{\lambda}\!:=\sum_{m=1}^{\infty}(-1)^{m+1}
\dim {\widetilde V}(\lambda)_{m\beta_l}$
linear exotic factors of $\det PRV^{\lambda}$. By~(\ref{estimprvex})
the number of linear
exotic factors is at most $l\cdot m_{\lambda}$. In order to obtain
the rest of the factors we are going to prove in this subsection
the following democracy theorem establishing
"equality of the odd positive roots
according Law".
\begin{thm}{thmdem}
(i) For any $\lambda\in P^+(\pi)$ and any $\mu\in {\frak h}^*$ such that
$(\mu+\rho,\beta_i)=0$ one has
$$\corank PRV^{\lambda}(\mu)\geq \sum_{m=1}^{\infty}(-1)^{m+1}
\dim {\widetilde V}(\lambda)_{m\beta_l}.$$
(ii) For any $\lambda\in P^+(\pi)$, $\beta_i\in {\Delta_1^+}$
the determinant $\det PRV^{\lambda}$ is divisible by
$$\left(\varphi(\beta_i)+(\rho,\beta_i)\right)^{\sum_{m=1}^{\infty}(-1)^{m+1}
\dim {\widetilde V}(\lambda)_{m\beta_l}}.$$
\end{thm}
\subsubsection{}
Let $s_i\in W$ be the simple reflection with respect to the simple root
$\alpha_i$,
where $\alpha_i=(\beta_i-\beta_{i+1})$ for $i<l$ and $\alpha_l=\beta_l$.
In order to prove the theorem above we need the following lemma.
\begin{lem}{estimN}
Fix $i\in\{1,\ldots,l-1\}$ and $n\in {\Bbb N}^+$. Take $\mu\in {\frak h}^*$
such that
$$
\begin{array}{c}
(\beta_i,\mu+\rho)=n,\ (\beta_{i+1},\mu+\rho)=0,\\
\text {the collection}\{(\beta_k,\mu+\rho)\}_{ k\not=i+1}\
\text { is linearly independent over }{\Bbb Q}.
\end{array}
$$
(i) Let $N$ be the kernel of the surjective map ${\widetilde M}(\mu)\to
{\widetilde V}(\mu)$.
Then
$$\underset{\xi\in\Omega (N)}\min (\mu-\xi,\omega_i)=n.$$
(ii) Assume that $\lambda\in P^+(\pi)$ is such that $(\lambda,\omega_i)<n$.
Then
$$\overline{C}(\lambda)\cap \Ann {\widetilde V}(\mu)\subseteq
\overline{C}(\lambda)\cap \Ann {\widetilde V}(s_i.\mu).$$
where $\overline{C}(\lambda)$ is the isotypical component of
$\widetilde{V}(\lambda)$ in ${\cal U}({\frak g})$.
\end{lem}
\begin{pf}
(i) The assumption on $\mu$ implies that
$\langle\mu+\rho,\alpha\rangle\in {\Bbb N}^+$ only when
$\alpha\!=\!\beta_i\pm\beta_{i+1}$ or $\alpha\!=\!\beta_{i}$.
Since $\langle\mu+\rho,\beta_{i}\rangle=2n\not\in 2{\Bbb N}+1$
and $\langle\mu+\rho,\beta_{i}\pm\beta_{i+1}\rangle=n$, we
conclude from~\ref{weightsub} that
$$\mu-\Omega (N)=\{n(\beta_i-\beta_{i+1})+{\Bbb N}(\pi)\}\cup
\{n(\beta_i+\beta_{i+1})+{\Bbb N}(\pi)\}.$$
Hence
$$\underset{\xi\in\Omega (N)}\min (\mu-\xi,\omega_i)=
\underset{\nu\in {\Bbb N}(\pi)}\min \{(n(\beta_i-\beta_{i+1})+\nu,\omega_i),
(n(\beta_i+\beta_{i+1})+\nu,\omega_i)\}=n$$
as required.
(ii) Since $(\mu+\rho,\beta_i-\beta_{i+1})\in {\Bbb N}^+$ it follows that
${\widetilde V}(s_i.\mu)$ is a subquotient of ${\widetilde M}(\mu)$. Therefore
$\Ann {\widetilde M}(\mu)\subseteq \Ann {\widetilde V}(s_{\alpha_i}.\mu)$.
Thus it is sufficient to show that
$$\overline{C}(\lambda)\cap \Ann {\widetilde V}(\mu)=
\overline{C}(\lambda)\cap \Ann {\widetilde M}(\mu).$$
Assume the contrary, namely, that there exists an
$\ad {\frak g}$-submodule $V\cong {\widetilde V}(\lambda)$ of
$\Ann {\widetilde V}(\mu)$ such that $V{\widetilde M}(\mu)\not=0$.
Since $V$ is $\ad {\frak g}$-invariant one has
$V{\cal U}({\frak g})={\cal U}({\frak g})V$. Consequently,
$V{\widetilde M}(\mu)\not=0$ forces $Vv_{\mu}\not=0$, where
$v_{\mu}$ is a highest weight vector of ${\widetilde M}(\mu)$.
Recall from~\ref{duality} that $\Omega (V)=-\Omega (V)$.
Therefore
\begin{equation}\label{omegan}
\underset{\xi\in\Omega (Vv_{\mu})}\max (\mu-\xi,\omega_i)=
\underset{\nu\in\Omega (V)}\max (-\nu,\omega_i)=
\underset{\nu\in\Omega (V)}\max (\nu,\omega_i)<n.
\end{equation}
On the other hand, $V{\widetilde V}(\mu)=0$ and so $Vv_{\mu}\subseteq N$.
Thus~(\ref{omegan}) contradicts to (i).
\end{pf}
\subsubsection{Proof of~\Thm{thmdem}}
Fix $\lambda\in P^+(\pi)$ and set
$$r=\sum_{m=1}^{\infty}(-1)^{m+1}\dim {\widetilde V}(\lambda)_{m\beta_l}.$$
For $k=1,\ldots,l$ let $x_k$ be the element of ${\cal S}({\frak h})$ given by
the formula $x_k\!:=\varphi(\beta_k)+(\rho,\beta_k)$.
Recall that $s_i$, $i=1,\ldots,l-1$, is the simple reflection with respect to
the root $(\beta_i-\beta_{i+1})$ and consequently one has, for any
$i=1,\ldots,l-1$
\begin{equation}
\label{xksi}
x_i(s_i.\mu)=x_{i+1}(\mu),\ x_{i+1}(s_i.\mu)=x_i(\mu),\ \
x_k(s_i.\mu)=x_k(\mu) \text { for } k\not=i,i+1.
\end{equation}
Each entry $a_{i,j}$ of the matrix $PRV^{\lambda}$ is an element
of ${\cal S}({\frak h})$ and so can be considered
as a polynomial in $x_1,\ldots , x_l$.
We have to prove that for any $i=1,\ldots,l$ and for any $\mu$ such that
$x_i(\mu)=0$, the corank of the matrix
$PRV^{\lambda}(x_1(\mu),\ldots ,x_l(\mu))$ is at least $r$.
We prove this assertion by induction on $i$.
By~\Cor{corbeta}, the assertion holds for $i=l$.
Assume that the assertion holds for some $i+1:\ 1<i<l$ and deduce it for $i$.
Recall~\Claim{clprv}:
\begin{equation}
\label{anncorank}
[\Ann_{{\cal H}} {\widetilde V}(\mu):{\widetilde V}(\lambda)]=
\corank PRV^{\lambda}(\mu)
\end{equation}
Thus our assumption implies that for any $\mu$ such that
$(\mu+\rho,\beta_{i+1})=0$ one has
\begin{equation}
\label{annsi}
[\Ann_{{\cal H}} {\widetilde V}(\mu):{\widetilde V}(\lambda)]\geq r.
\end{equation}
Fix, for a moment, a collection of complex numbers $\{a_k\}_{k=1}^{l}$
such that
the collection $\{1,a_k:k=1,\ldots,l\}$ is linearly independent over
${\Bbb Q}$.
For every $n\in {\Bbb N}^+$, $n>(\lambda,\omega_i)$, take
$\mu_n\in {\frak h}^*$
such that $x_i(\mu_n)=n,\ x_{i+1}(\mu_n)=0,\ x_k(\mu_n)=a_k$ for
$k\not=i,i+1$.
Then~\Lem{estimN} (ii) combined with the inequality~(\ref{annsi}) implies that
\begin{eqnarray*}
r\leq [\Ann_{{\cal H}} {\widetilde V}(s_i.\mu_n):{\widetilde V}(\lambda)]&=&
\corank PRV^{\lambda}(s_i.\mu_n)\\
&=&\corank PRV^{\lambda}(x_1(s_i.\mu_n),\ldots ,x_l(s_i.\mu_n)).
\end{eqnarray*}
Thus $\corank PRV^{\lambda}$ is greater than or equal to $r$ at the points
$$(x_1(s_i.\mu_n),\ldots,x_l(s_i.\mu_n))=
(a_1,\ldots,a_{i-1},0,n,a_{i+2},\ldots, a_l)\ \ \mbox{ (see~(\ref{xksi}))}$$
of the line $$(a_1,\ldots,a_{i-1},0,{\Bbb C},a_{i+2},\ldots, a_l).$$
Therefore $\corank PRV^{\lambda}\geq r$ on the whole line
$(a_1,\ldots,a_{i-1},0,{\Bbb C},a_{i+2},\ldots, a_l)$.
Observe that the union of the lines
$(a_1,\ldots,a_{i-1},0,{\Bbb C},a_{i+2},\ldots, a_l)$ such
that $\{1,a_k:k=1,\ldots,l\}$ is linearly independent over ${\Bbb Q}$ is dense
in the hyperplane $x_i=0$. Hence $\corank PRV^{\lambda}$
is greater than or equal to
$r$ in the whole hyperplane $x_i=0$ as required.
This establishes~\Thm{thmdem}.
\subsection{}
\label{endpfprv}
Combining the results of
~\Cor{stfact},~\Thm{thmdem} and the bound~(\ref{estimprvex})
we obtain~\Thm{thmprv}.
\section{Main theorem}
\subsection{}
\begin{thm}{main} Take $\lambda\in {\frak h}^*$. Then
\begin{equation}
\label{maineq}
\Ann_{{\cal U}({\frak g})}{\widetilde M}(\lambda)=
{\cal U}({\frak g})\Ann_{{\cal Z}({\frak g})}{\widetilde M}(\lambda)\ \
\Longleftrightarrow \ \ \forall\alpha\in \Delta_1^+:\
(\lambda+\rho,\alpha)\not=0.
\end{equation}
\end{thm}
The standard factors of~\Thm{thmprv} are exactly the factors which appear in
the factorization of Shapovalov determinants---see~\Thm{shapovalov}.
Since a Verma module $\widetilde{M}(\mu)$ is simple iff $\mu$ is
not a root of any Shapovalov determinant, this proves the
equivalence~(\ref{maineq}) for simple Verma modules. By~\ref{redtosimple},
it gives immediatly the implication "$\Leftarrow$" of~(\ref{maineq}).
\subsection{}
\label{otherim}
Let us prove the other implication.
Consider ${\frak g}_1$ as a ${\frak g_0}$-module. All its
weight-spaces are one-dimensional,
$\Omega ({\frak g}_1)=\{\pm\beta_1,\ldots,\pm\beta_l\}$ so
${\frak g}_1\cong V(\omega_1)$. Consider the
natural representation $\widetilde{V}(\omega_1)$ of $\osp(1, 2l)$.
It is easy to see that
$\widetilde{V}(\omega_1)=V(\omega_1)\oplus V(0)$ as ${\frak g_0}$-modules.
Again all weight-spaces of $\widetilde{V}(\omega_1)$ are one-dimensional
and $\Omega (\widetilde{V}(\omega_1))=\{0,\pm\beta_1,\ldots,\pm\beta_l\}$.
Hence, by~\Thm{thmhess},
${\cal H}$ contains exactly one copy $V$ of $\widetilde{V}(\omega_1)$
which occurs in degree $2l$.
Fix $i\in\{1,\ldots,l\}$. One has
$$\mathop{\sum}_{m=1}^{\infty}{(-1)}^{m+1}
\dim{\widetilde{V}(\omega_1)}_{m\beta_i}=1$$
and so, by~\Thm{thmprv}, $\det PRV^{\omega_1}(\mu)=0$ for any
$\mu\in {\frak h}^*$
such that $(\mu+\rho,\beta_i)=0$. By~\Lem{clprv}, it implies that
$V\subset \Ann_{{\cal U}({\frak g})}\widetilde{V}(\mu)$.
In other words, $V\subset \Ann_{{\cal U}({\frak g})}\widetilde{M}(\mu)$
for any $\mu$ in the set
$$\left\{\mu\in {\frak h}^*|\ (\mu+\rho,\beta_i)=0 \mbox { and }
\widetilde{M}(\mu)\mbox{ is simple}\right\}.$$
According to the criterion of simplicity given in~\Cor{simpleV}, this set is
Zariski dense in the hyperplane
$(\mu+\rho,\beta_i)=0$.
The following lemma proves that in fact
$V\subset \Ann_{{\cal U}({\frak g})}\widetilde{M}(\mu)$
for any $\mu$ such that $(\mu+\rho,\beta_i)=0$.
This gives the implication "$\Longrightarrow$" of~(\ref{maineq}).
\subsection{}
\begin{lem}{zarcl}
Let $V$ be an $\ad$-invariant subspace of ${\cal U}({\frak g})$. Then the set
$\left\{\lambda\in{\frak h}^*,\ V\subset\Ann_{{\cal U}({\frak g})}
\widetilde{M}(\lambda)\right\}$ is a Zariski closed subset of ${\frak h}^*$.
\end{lem}{}
\begin{pf} For any $\lambda\in {\frak h}^*$ denote by $v_{\lambda}$
a highest weight vector of $\widetilde{M}(\lambda)$.
Remark that the $\ad$-invariance of $V$ implies that
${\cal U}({\frak g})V=V{\cal U}({\frak g})$.
Hence, for any $\lambda\in {\frak h}^*$, one has the equivalence $V\subset
\Ann_{{\cal U}({\frak g})}\widetilde{M}(\lambda)
\Longleftrightarrow V.v_{\lambda}=0$. Let
$\{a_j\}_{j\in J}$ be a basis of $V$. Then
$$\left\{\lambda\in{\frak h}^*,\ V\subset\Ann_{{\cal U}({\frak g})}
\widetilde{M}(\lambda)\right\}=\mathop{\bigcap}_{j\in J}
\left\{\lambda\in {\frak h}^*,\ x.v_{\lambda}=0\right\}.$$
Thus it is sufficient to show that for every $a\in {\cal U}({\frak g})$,
the set $\left\{\lambda\in {\frak h}^*,\ a.v_{\lambda}=0\right\}$
is a Zariski closed subset of ${\frak h}^*$.
Using the triangular decomposition ${\cal U}({\frak g})=
{\cal U}({\frak n}^-)\otimes{\cal U}({\frak h})\otimes{\cal U}({\frak n}^+)$,
write
$a=y_1P_1+\ldots+y_rP_r+x$ where $y_1,\ldots,y_r$ are linearly independent
elements of ${\cal U}({\frak n}^-)$, $P_1,\ldots,P_r\in{\cal U}({\frak h})$
and $x\in {\cal U}({\frak g}){\frak n}^+$. Then
$$\begin{array}{lcl} a.v_{\lambda}=0 &\Longleftrightarrow & \bigl(\displaystyle
\mathop{\sum}_i y_iP_i\bigr).v_{\lambda}=0\\
&\Longleftrightarrow & \displaystyle\mathop{\sum}_i P_i(\lambda)y_i.
v_{\lambda}=0\\
&\Longleftrightarrow & \forall i\ P_i(\lambda)=0.\
\end{array}.$$
The lemma is proved.
\end{pf}
This completes the proof of~\Thm{main}.
|
\section{Introduction}
The ATLAS Collaboration has carried out a detailed study to detect the
SUSY signature in the framework of one of the most popular model,
SUGRA~\cite{SUGRA_rev},\cite{SUGRA_pap}.
It has been shown~\cite{SUGRAincl} that if SUSY exists
at the electro-weak scale, it should be discovered by ATLAS and a general
method has been given to determine approximately the mass scale of the
SUSY particles. In subsequent papers~\cite{SUGRA12}--\cite{SUGRA5}
it was shown in five representative points of the parameter space
that some of the SUSY particles can be reconstructed and using the obtained
characteristics (masses, branching ratios) the model parameters can be
precisely determined \cite{Froid_LHCC}. All these studies have been carried out
assuming that $R$ parity is conserved.
In this note we consider that $R$ parity is broken in such a way that
the lepton number $L$ is violated through $\lambda$-type couplings.
The present experimental limits~\cite{Rlimits}
cannot completely exclude such a scenario. In this case one of the prominent
signatures of SUSY, the missing energy is considerably weakened because
the lightest SUSY particle (LSP) is allowed to decay. Due to this decay
the lepton and/or jet multiplicity increases considerably and some efficient
cuts (e.g. lepton veto against the $t-\bar t$ background) cannot be applied.
On the other hand,
the decay products of the LSP in some cases allow its direct reconstruction.
Therefore the event topology and the search strategies are different of the
case when $R$ is conserved. This has motivated us to revisit
the feasability to detect SUSY and to determine the parameters of the SUGRA
model using the ATLAS detector.
In section 2 we give a brief description of the phenomenology of the $R$ parity
violation and the event generator used.
Section 3 deals with the ATLAS detector and with the fast simulation
of its response. In section 4 we present the domain of the parameter space where
a SUSY signal can be expected by ATLAS. In the subsequent three sections the
reconstruction of the SUSY particles and the determination of the model parameters
are described in the LHC points 1,3 and 5 which represent a heavy, light and
medium SUSY mass scale. We summarize the obtained results
in the concluding section.
\vskip 1.2 cm
\section{Basic Phenomenology}
\vskip 0.5 cm
\subsection{$R$ parity violation}
\vskip 0.5 cm
$R$-parity has been introduced~\cite{Fayet1} in order to avoid fast nucleon decay
and flavor changing neutral currents (FCNC). If the multiplicative quantum number
\begin{equation}
R = (-1)^{3B+L+2S}
\label{eq:Rdef}
\end{equation}
is conserved it guarantees automatically baryon number ($B$) and lepton number
($L$) conservation. $R$ is +1 for Standard Model (SM) particles and its value is
-1 for their superpartners. The most important experimental consequences of the
conservation of $R$ are that super partners should be produced in pairs and the
lightest superpartner (LSP) should be stable. The LSP interacts weakly, therefore
the prominent signature of SUSY in case of $R$ parity conservation is a considerable
amount of missing (transverse) energy ($E_T^{miss}$).
Although no violation of $B$ or $L$ has been observed
yet, there is no firm theoretical argument which would require exact conservation
of them and that of the $R$ parity. In fact the following term in the superpotential
\begin{equation}
W_{\rlap/R} = \lambda_{ijk}L_iL_jE_k^c + \lambda_{ijk}'Q_iL_jD_k^c +
\lambda_{ijk}''U^c_iD^c_jD_k^c
\label{eq:Lag}
\end{equation}
which violates explicitely $B$, $L$ and $R$ parity, cannot be ruled out experimentally.
Here $L$ and $E$ are isodublet and isosinglet lepton, $Q$ and $D$ are isodublet and
isosinglet quark superfields, the indices $i$, $j$ and $k$ run for the three lepton
and quark families. The suffix $c$ denotes charge conjugate.
The first two terms violate explicitely $L$ whilst the last
one violates $B$. Present limits on the proton lifetime suggests that either the
$L$ or the $B$ violating terms (i.e. the corresponding $ \lambda_{ijk}$ couplings)
should vanish for the first family. Other experimental limits e.g. on lepton number violation:
double $\beta$ decay, or on $N-\bar N$ oscillation, etc. indicate that the couplings
in Equ.~(\ref{eq:Lag}) shouldn't be expected to exceed a few percent, and usually are
much smaller than the gauge couplings.
Even so, if $R$ parity is violated the topology of the expected SUSY signal changes
substantially. Since the LSP is no more stable, the missing energy is
considerably reduced. On the other hand the decay products of the LSP increase
the average number of jets and/or leptons in an event. In general, the event topology
depends crucially on the size of the couplings. If, e.g. the couplings are
of the order of $\sim 10^{-2}$ or larger, the mass spectra, branching ratios, etc.
will be different in the two cases where $R$ is conserved or violated.
If however the couplings are smaller than the above value, the
dominant effect of the $R$ parity violation is that the LSP becomes unstable.
An estimation of the LSP lifetime as a function of the couplings show~\cite{lifetime}
that we can distinguish four subcases giving rise to different detection
strategies: \\
\indent \indent \indent $(i)$ $10^{-4} \leq \lambda \leq 10^{-2}$ \\
\indent \indent \indent $(ii)$ $10^{-6} \leq \lambda \leq 10^{-4}$ \\
\indent \indent \indent $(iii)$ $10^{-9} \leq \lambda \leq 10^{-6}$ \\
\indent \indent \indent $(iv)$ $\lambda \leq 10^{-9}$ \\
In case $(i)$ only the event topology changes w.r.t. the case of $R$ conservation.
In case $(ii)$ one can observe a displaced vertex at the LHC energies.
In case $(iii)$ the LSP decays outside of a typical LHC detector, however
it can be catched by special purpose detectors~\cite{LSPdecay}. Finally
the case $(iv)$ cannot be distinguished experimentally from the case of $R$ parity
conservation.
In this study we have deliberately chosen to study case $(i)$ and compare
the result with the case of $R$ parity conservation, because it represents a more
difficult experimental situation than case $(ii)$ where a displaced vertex could
disentangle the LSP from the rest of the event. Moreover we have assumed that
$\lambda_{ijk}' = \lambda_{ijk}'' = 0$ and only
one of the $\lambda_{ijk}$ coupling is different from zero in Equ.~(\ref{eq:Lag}).
Nonzero $\lambda_{ijk}'$ and
$\lambda_{ijk}''$ are subject of other reports inside ATLAS~\cite{Jesper}.
The hierarchical structure observed for the Yukawa couplings in the
SM motivates our hypothesis above. The Lagrangian corresponding to
the superpotential of Equ.~(\ref{eq:Lag}) can be written
in terms of particle fields for our case as:
\begin{equation}
L^{\lambda}_{\not L} = \frac{1}{2}\lambda_{ijk}
(\overline{\nu}^{c}_{L_{i}} e_{L_{j}} \tilde e^{*}_{R_{k}} -
e_{L_{i}} \overline{\nu}^{c}_{L_{j}} \tilde e^{*}_{R_{k}} +
\nu_{L_{i}} \tilde e_{L_{j}} \overline{e}_{R_{k}} -
e_{L_{i}} \tilde \nu_{L_{j}} \overline{e}_{R_{k}} +
\tilde \nu_{L_{i}} e_{L_{j}} \overline{e}_{R_{k}} -
\tilde e_{L_{i}} \nu_{L_{j}} \overline{e}_{R_{k}}) + HC
\label{eq:Lagl}
\end{equation}
where $\nu_{L}$ and $e_{L,R}$ are lepton fields, the tilde
denotes the field of the superpartner, the $c$ stands for charge conjugation,
$*$ for complex conjugation and $i,j,k$ are the flavor indices.
As stated before, if the $\lambda$ couplings are smaller than $10^{-2}$,
which is our case, the sparticle mass spectrum
practically doesn't change and the main consequence of the $R$ parity violation is
the decay of the LSP. This process is depicted in Fig.\ref{LSPdecay1}
where we assume that the LSP is the lightest neutralino ($\tilde \chi^0_1$).
The decay proceeds through an $R$ conserving and an $R$ violating vertex
and in the final state there are always three leptons out of them at least
two of {\it different}
flavours, one neutral and the other two of opposite charges, since the
LSP is supposed to be neutral.
The prominent signature of this type of SUSY event is the spectacular
increase of "stable" leptons: electrons and muons in the final state.
The neutral leptons, neutrinos, give rise to some missing transverse
energy, but its magnitude is much less than that if $R$ parity is conserved.
The flavour of the lepton in the final state depends on the values
of the indices $i,j,k$. Since $\lambda_{ijk}$ is antisymmetric in $i$ and
$j$ there are only 9 independent couplings which we choose as:
$\lambda_{121},\lambda_{122},\lambda_{123},\lambda_{131},\lambda_{132},
\lambda_{133},\lambda_{231},\lambda_{232},\lambda_{233}$. The first two
families, 1 and 2 give rise always to "stable" leptons, electrons and
muons, (and the corresponding neutrinos) in the final state. If
an index 3 appears, the lepton is not stable if it is a $\tau$,
and its decay products are most of the time different from electrons
or muons. Since $e$ in Equ.~(\ref{eq:Lag}) is an isosinglet, $e_3$
is a $\tau$. The number of the stable leptons is the most
prominent for $\lambda_{121}$ and $\lambda_{122}$. It is less spectacular,
if an index 3 appears at the second place, and even less if the 3 appears
in the third place. Finally, if two indices have value of 3 one has the least
number of stable leptons. On the other hand, the number of the
neutrinos, and with that the magnitude of the missing energy
increases in the order of the above mentioned cases.
The expected extra number of stable charged leptons
and neutrinos per each LSP decay for the different $\lambda$ couplings calculated
by the program \cite{isajetR} are given in
Table~\ref{tb:BRs}. It is clear that the average number
of the leptons for different flavours gives a strong hint
on the coupling which is realized. E.g. the coupling $\lambda_{122}$
gives rise predominantly to muons whilst the coupling $\lambda_{123}$
results in equal number of electrons and muons, etc.
\vskip -0.5 cm
\begin{Fighere}
\centering
\epsfig{file=./feyn.eps,width=14cm }
\vskip -10.0 cm
\caption{\small The $\rlap/R$ decays of $\chi^{0}_{1}$ (assumed as $LSP$) through
$\lambda$ type couplings.}
\label{LSPdecay1}
\end{Fighere}
\vskip0.5cm
\vspace{-0.5cm}
\begin{Tabhere}
\centering
\caption{\small The number of produced stable leptons in the LSP decay
if it is the $\tilde \chi^0_1$. We have assumed $\sim$ 18\% branching
ratio for the semileptonic decay of the $\tau$.}
\vskip 0.5cm
\begin{tabular}{|c|c|c|c|c|} \hline
$\lambda_{ijk}$ & Decay channel & $<N_e>$ & $<N_{\mu}>$ & $<N_{\nu}>$ \\ \hline
121 & $e^{\pm}\nu_{\mu}e^{\mp}$;$\nu_e\mu^{\pm}e^{\mp}$ & 1.5 & 0.5 & 1 \\ \hline
122 & $e^{\pm}\nu_{\mu}\mu^{\mp}$;$\nu_e\mu^{\pm}\mu^{\mp}$ & 0.5 & 1.5 & 1 \\ \hline
123 & $e^{\pm}\nu_{\mu}\tau^{\mp}$;$\nu_e\mu^{\pm}\tau^{\mp}$ & 0.68 & 0.68 & 2.36 \\ \hline
131 & $e^{\pm}\nu_{\tau}e^{\mp}$;$\nu_e\tau^{\pm}e^{\mp}$ & 1.59 & 0.09 & 1.68 \\ \hline
132 & $e^{\pm}\nu_{\tau}\mu^{\mp}$;$\nu_e\tau^{\pm}\mu^{\mp}$ & 0.59 & 1.09 & 1.68 \\ \hline
133 & $e^{\pm}\nu_{\tau}\tau^{\mp}$;$\nu_e\tau^{\pm}\tau^{\mp}$ & 0.77 & 0.27 & 3.04 \\ \hline
231 & $\mu^{\pm}\nu_{\tau}e^{\mp}$;$\nu_{\mu}\tau^{\pm}e^{\mp}$ & 1.09 & 0.59 & 1.68 \\ \hline
232 & $\mu^{\pm}\nu_{\tau}\mu^{\mp}$;$\nu_{\mu}\tau^{\pm}\mu^{\mp}$ & 0.09 & 1.59 & 1.68 \\ \hline
233 & $\mu^{\pm}\nu_{\tau}\tau^{\mp}$;$\nu_{\mu}\tau^{\pm}\tau^{\mp}$ & 0.27 & 0.77 & 3.04 \\ \hline
\end{tabular}
\label{tb:BRs}
\end{Tabhere}
\vskip 0.5cm
The presence or absence of $\tau$'s influence how easily the
SUSY particles can be reconstructed. On the one hand the signature
is weakened by the taus since they produce less stable charged
leptons, and these, partly originating from the $\tau$ decay,
do not produce a sharp endpoint for the $\tilde\chi_1^0$ mass. On
the other hand, taus produce more missing transverse energy,
moreover the smaller number of stable charged leptons produced
decreases the combinatorial background in the mass reconstruction.
\subsection{The SUGRA Model}
The present study of $R$ parity violation is carried out
in the framework of the SUGRA model~\cite{SUGRA_rev},\cite{SUGRA_pap}.
Contrary to the minimal version of supersymetric models (MSSM)
which has a very large number of unknown parameters, the SUGRA model
is characterized only by 5 parameters which are the following: \\
\indent $(1)$ $m_0$, an universal scalar mass, \\
\indent $(2)$ $m_{1/2}$, an universal gaugino mass, \\
\indent $(3)$ $A_{0}$ a common trilinear interaction term, \\
\indent $(4)$ tan($\beta$), the ratio of the vacuum expectation values
of the two Higgs fields,\\
\indent $(5)$ the sign of $\mu$ of the Higgsino mass parameter.\\
The mass spectrum of the SYSY partners at the
electro-weak scale as well as their decay branching ratios
are obtained from the above parameters
by solving the renormalization group equations (RGE). This is
performed in our case by the program of ISAJET~\cite{isajet}.
The SUGRA model predicts a hierarchical structure of the masses
of the SUSY particles.
The masses of the first two families of the squarks and of
the sleptons are driven essentially by
$m_{0}$ and $m_{1/2}$ through an approximate relation :
\begin{equation}
m^{2}_{\tilde f_{L,R}} = m^{2}_{0}+m^{2}_{f}+c(\tilde f_{L,R}) \cdot m^{2}_{1/2}+
D(\tilde f_{L,R})
\label{eq:squarkmasses}
\end{equation}
where $c(\tilde f)$ are some numerical factors of order $5.5 \div 6.0$ for the
squarks and of order
$0.15 \div 0.5$ for the sleptons, $D(\tilde f_{L,R})$ are the so-called D-terms (in general less
important). For the third family of the squarks and the sleptons
there is a mixing due to the
corresponding Yukawa couplings, pushing down the mass of one of the sfermions
and the other one in the opposite
direction. In some regions of the parameter space this could have the effect that the lighter
stau ($\tilde \tau_{1}$) could become the LSP but such scenarios are ruled out by cosmological
considerations. The gauginos $U(1),SU(2)$ and $SU(3)$ are driven mainly by $m_{1/2}$.
Because the $U(1)$ and $SU(2)$ gauginos will mix with higgsinos to obtain the mass states
($\tilde \chi^{0}_{i}$, $i=1,4$,
and $\tilde \chi^{\pm}_{j}$, $j=1,2$), part of that spectrum will depend on the $\mu$ parameter
(in SUGRA $\mu$ is determined by the condition of electro-weak symmetry breaking and typically
$\mu \gg m_{1/2}$).
We have the following approximate relations:
\vskip 1.0 cm
\begin{equation}
\centering
\begin{tabular}{c}
$m_{\tilde g} \approx 2.4 \cdot m_{1/2}$ \\
$m_{\tilde \chi^{0}_{1}} \approx 0.4 \cdot m_{1/2}$ \\
$m_{\tilde \chi^{0}_{2}} \approx m_{\tilde \chi^{\pm}_{1}} \approx 0.8 \cdot m_{1/2}$ \\
$m_{\tilde \chi^{0}_{3}} \approx m_{\tilde \chi^{0}_{4}} \approx m_{\tilde \chi^{\pm}_{2}} \approx |\mu|$ \\
\end{tabular}
\label{eq:gauginomasses}
\end{equation}
\vskip 1.0 cm
The Higgs sector is composed in SUGRA by five mass states ($h^{0}$,
$A^{0}$, $H^{0}$ and $H^{\pm}$).
In the case of the lightest one, $m_{h^0} \leq 130 \div 150$ GeV
in about all SUGRA parameter space.
The masses of the other Higgses are in general
very heavy and depend on tan$( \beta )$.
The gaugino-higgsino mixing depends more strongly on the parameter
values, and this in turn determines the decay branching ratios.
\newpage
\section{The analysis chain}
\subsection{The event generators}
\indent \indent {\it \bf Simulation of the signal events}
\vskip 0.2cm
Our basic program tool is ISAJET\footnote{We have used version 7.30}
\cite{isajet} which simulates
for hadron colliders the production and decay of the supersymetric
particles, as well as the underlying event, i.e. the accompanying
partons and their hadronization.
ISAJET does not include $R$ parity violation
in the SUGRA model. It has been introduced in detail in another
event generator, SUSYGEN \cite{SUSYGEN}, written for $e^+e^-$
collisions.
Therefore, inspired by SUSYGEN,
a set of routines computing in detail all the $\rlap/R$ ($\lambda$-type) decays
of gauginos ($\tilde \chi^{0}_{i}$, $i=1,4$ and $\tilde \chi^{\pm}_{j}$, $j=1,2$) and
sleptons was written for ISAJET~\cite{isajetR}. This program is able to simulate production
of supersymetric particles in a hadronic collision with $R$ parity broken, if this latter
is accompanied by $L$ number violation.
To obtain correct predictions, the hypothesis of small values of the $\lambda$ couplings
($\lambda \leq 10^{-2}$) must be used, which is our present case,
in order to neglect any correction in the sparticle
mass spectrum brought by RGE's.
The dominant effect of the $R$ parity violation is the decay of the LSP.
We have generated several million signal events. In Table~\ref{tb:sigstat}
one can see their repartition in the SUGRA parameter space and for the
type of the couplings. In the last column we quote the integrated luminosity
in terms of LHC months the generated events correspond to where we have
taken an LHC year equivalent to $10^4$ /pb (low lumonosity run with
appr. 1/3 of efficiency).
\vspace{-0.5cm}
\begin{Tabhere}
\centering
\caption{\small The number of generated signal events }
\vskip 0.5cm
\begin{tabular}{|l|c|c|c|c|c|r|c|} \hline
$m_0$ [GeV] & $m_{1/2}$ [GeV] & $A_0$ & tan($\beta$) & sgn($\mu$) & coupling & Events & LHC month \\ \hline
0 - 1500 & 0 - 1500 & 0 & 2 & +1 & $\lambda_{123}$ & 256 000 & --- \\ \hline
0 - 500 & 0 - 500 & 0 & 2 & +1 & $\lambda_{123}$ & 300 000 & --- \\ \hline
400 & 400 & 0 & 2 & +1 & $\lambda_{122}$ & 400 000 & $\sim 36$ \\ \cline{6-8}
& & & & & $\lambda_{123}$ & 400 000 & $\sim 36$ \\ \hline
200 & 100 & 0 & 2 & -1 & $\lambda_{123}$ & 1 000 000 & $\sim 1.5$ \\ \cline{6-8}
& & & & & $\lambda_{122}$ & 1 000 000 & $\sim 1.5$ \\ \hline
100 & 300 & 300 & 2.1 & +1 & $\lambda_{123}$ & 450 000 & $\sim 36$ \\ \cline{6-8}
& & & & & $\lambda_{122}$ & 450 000 & $\sim 36$ \\ \hline
\end{tabular}
\label{tb:sigstat}
\end{Tabhere}
\vskip 0.5cm
The samples in the first two raws were used to study inclusive reactions. The rest of the statistics is devoted
to the 1st, 3rd and 5th of the so called LHC points, where the determination of the SUGRA parameters
has been performed.
\vskip 0.2cm
{\it \bf Simulation of the background events}
\vskip 0.2cm
As mentioned, the signature of the signal events is the appearence of a large number
of high $p_t$ leptons accompanied by high $p_t$ jets and by a moderate amount of
missing transverse energy. Such event topology is also produced, albeit in a much
reduced level, by the decay of heavy SM particles and these events constitute the
main background. We have studied such processes using the ISAJET and PYTHIA~\cite{pythia}
event generators including initial and final state radiation.
The background event statistics is listed in Table\ref{tb:bgstat}.
\vspace{-0.5cm}
\begin{Tabhere}
\centering
\caption{\small The number of generated background events }
\vskip 0.5cm
\begin{tabular}{|l|c|} \hline
Reaction & Number of events \\ \hline
$t-\bar t$ & 1 200 000 \\ \hline
$W-Z$, $W-W$ and $Z-Z$ & 600 000 \\ \hline
$Z-b\bar b$ & 600 000 \\ \hline
Drell-Yan & 600 000 \\ \hline
\end{tabular}
\label{tb:bgstat}
\end{Tabhere}
\vskip 0.5cm
\subsection{Fast simulation of the ATLAS detector}
Due to the large number of events to be generated the detector response could not
have been simulated in detail, using e.g. GEANT~\cite{geant}. Instead a fast,
so called particle level Monte Carlo program, ATLFAST~\cite{atlfast}
has been used\footnote{We have used the version 1.57}.
In this program the ATLAS detector~\cite{ATLAS} is described by a simplified geometry
and apart of the acceptance the detector response is parametrized. Below we repeat
the main features of these description.
\vskip 0.2cm
{\it \bf Description of ATLAS by ATLFAST}
\vskip 0.2cm
The detector geometry is given in the variables of the pseudorapidity
$\eta = - \mbox{ln tan}(\theta /2)$ and azymuthal angle $\phi$, where
$\theta$ is the solid angle of the particle produced in the interaction
point. The granularity in $\eta \times \phi$ is 0.1 $\times$ 0.1 for
$|\eta | < 3$ and 0.2 $\times$ 0.2 otherwise upto $|\eta | = 5$.
The produced particles except the muons and the neutrinos
deposite their energies, smeared by a resolution function,
and are integrated in individual $\eta$-$\phi$ cells.
The effect of the 2T solenoidal magnetic field on the deposited energy
of the charged particles is parametrized. The effect of cracks
in the calorimeter and in the trackers are taken into account by
the parametrized acceptance function.
\vskip 0.2cm
{\it \bf Reconstruction algorithms}
\vskip 0.2cm
First clusters are created from the cell energies using a simple algorithm.
Next, isolated photons and electrons are reconstructed if
the simulated particle falls in the acceptance region
(typically $|\eta | < 2.5$), its energy,
smeared by a parametrized resolution function,
matches that of a cluster, and in the case of electrons, the $\eta$ and $\phi$
values of the cluster and the electron are the same within the resolution.
The energy of the reconstructed particles is that of the simulated one
smeared by a parametrized resolution, whereas their direction stays the same.
After having removed the clusters of the isolated photons and electrons
the reconstruction of the parton jets is carried out
by a simple fixed cone algorithm. Isolation criteria of jets
or charged tracks can be defined as a function of the deposited energy in a
$\Delta R = \sqrt{(\Delta\eta)^2+(\Delta\phi)^2}$ cone.
The reconstructed jet energy in general doesn't match with the
true energy of the corresponding parton.
Therefore a correction factor has to be applied in a later phase
of the analysis. This correction factor, depending on the
cone size and on the momentum
of the jet, is established by comparing the reconstructed jet transverse momentum
with that of the original parton: $R_{calib}^{b jet}= p_{t}^{b parton}/p_{t}^{b jet}$.
It has been demonstrated that applying such factors
one can correctly reconstruct the position of the mass peak
of the Higgs boson~\cite{atlfast}. This correction factor depends
on the jet type and also if the jet contains prompt leptons.
In Fig.\ref{p5selbjets} the correction is shown for the
three different LHC points analysed. The observed difference
in point 3 and the other two points is due to the facts that :\\
\indent $(i)$ the production mechanisms of the $b$'s are different. In point
1 and 5 they originate mainly from the decay $h^0\longrightarrow b\bar b$,
whereas in point 3 the production of the $b$'s is less correlated;\\
\indent $(ii)$ in point 3 there are more leptons produced and therefore
the probability is higher that a lepton is inside the jet-cone.\\
One has to stress that the dispersion of the correction factor
can be rather large, reaching even 25\%, especially at low $p_t$ ($\le $ 100 GeV).
Fig.\ref{p5res} shows the invariant mass of the $b \bar b$ pair from the $h^0$ mass
at the LHC points 1 and 5 after having applied the correction.
\vskip -0.5 cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./all_calib200.eps,height=10cm,width=14cm }}
\vskip -5.cm
\caption{\small Calibration functions of b jets for SUGRA points 1, 3 and 5. }
\label{p5selbjets}
\end{Fighere}
\vskip -0.5cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p1p5_resh0.eps,width=11cm }}
\vskip -5.0cm
\caption{\small
The effect of the resolution and calibration of b jets on the invariant mass distributions in
SUGRA points 1 (left) and 5 (right) and for the $\rlap/R$ couplings $\lambda_{122}=10^{-3}$:
$M(h^{0}) \sim M(bb)$ - for reconstructed b jets originating from $h^0$'s.
The ISR, FSR and hadronization are switched on. The arrows point to the theoretical values of
$m_{h^0}$ in each SUGRA point.}
\label{p5res}
\end{Fighere}
\vskip 0.5 cm
Muons are reconstructed if they are in the acceptance region
(typically $|\eta | < 2.5$) and their energies and directions
are obtained from the simulated one by a parametrized smearing
function of ATLFAST.
In most cases these functions were determined or
cross checked by detailed simulation using GEANT. Finally,
the missing transverse energy is calculated.
\subsection{Detection and identification efficiencies}
The output of ATLFAST is written in a coloumn-wise ntuple (CWN)
for further analysis using PAW~\cite{paw}. This ntuple contains
all the reconstructed electron, muon and jet objects together
with the two transverse components of the missing energy.
In addition, we have stored among the originally
simulated particles and partons by RPV\_ISAJET those which were produced in
the decay chain of the generated SUSY particles. We have installed
a bidirectional pointer between the reconstructed objects
and the original particles/partons, and we have reinstalled the mother-daughter
relationships between the stored original particles/partons.
A special record containing the integrated luminosity has been also included.
As a first step in the analysis we have randomly rejected reconstructed
electrons and muons using a tabulated detection (including identification)
efficiency. The efficiencies, as a function of the transverse
momentum and $\eta$, have been extracted from
a version of ATLFAST (2.0) which was released after the bulk of
our event simulation has been carried out (see Fig.\ref{effic}).
\vskip -0.5cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./id_lepton.eps,width=14cm }}
\caption{\small Detector efficiencies for muons and electrons
as a function of $p_t$ and $\eta$. The dependence
on $\varphi$ is practically negligible.}
\label{effic}
\end{Fighere}
\vskip 0.5 cm
We have also randomly
reclassified tagged b-jets to light jets with a b-tagging efficiency of 60\%
and vice versa using the misstaging probability of $\sim$2\%. These numbers
have been obtained in a separate study by detailed Monte Carlo
simulation~\cite{btagging}.
The obtained final state reconstructed particles were submitted
to selection cuts in order to find an optimum signal of SUSY
over the background. These criteria as well as the reconstruction
algorithms are described in the forthcoming sections.
\section{Inclusive measurements}
By measuring global variables, like e.g. the number
of leptons of a given flavour or the average $p_t$ of a lepton
in an event, or simply the number of events passing some
selection criteria, we would like to answer the following
three questions: \\
$(i)$ what is the maximum domain in the SUGRA parameter space
in which ATLAS is sensitive for $R$ parity violation; \\
$(ii)$ can we determine the approximate
energy scale of a SUGRA signal; \\
$(iii)$ can we determine the dominant type of the different couplings
which causes the signal.
For all these studies as well as for the exclusive measurements
we have fixed the values of the $\lambda_{ijk}$ couplings to $10^{-3}$.
We remind the reader that the event topology does not depend on the
particular value of $\lambda$ if it is between $10^{-2}$ and $10^{-4}$.
\vskip 0.2cm
{\it \bf Sensitivity of ATLAS in the SUGRA space}
\vskip 0.2cm
We have generated 1000 events in each of the 16$\times$16 equally
spaced points in the $m_{1/2}$ vs $m_0$ plane the other parameters
being fixed (see Table\ref{tb:sigstat}). We have generated the
SM background events as given in Table\ref{tb:bgstat}. Since the
signature of the SUGRA event is characterized by multileptons and
missing energy, we have tried several event selection based on
the variation of the number of electrons and muons, their transverse
energy and the value of the missing energy. We have used the
quantity called significance:
\begin{equation}
S = N_{sig}/\sqrt{N_{bg}}
\label{eq:significance}
\end{equation}
to delimite the sensitive region with $S \ge 5$, where $N_{sig}$
is the number of the accepted SUGRA events and $N_{bg}$
is the number of accepted SM events. The sensitive region
extending to the highest value of $m_0$ and $m_{1/2}$ has been
obtained using the selection criteria: \\
\indent \indent 1. $N_l \ge 3$ \\
\indent \indent 2. $E_t^{miss} \ge 100$ GeV \\
\indent \indent 3. $p_t^1 \ge 70$ GeV \\
\indent \indent 4. $p_t^l \ge 20$ GeV \\
where $N_l$ is the total number of $e$ and $\mu$,
$p_t^1$ is the highest transverse momentum of the leptons,
$p_t^l$ is the transverse momentum of the other leptons in the event.
The first two cuts reject almost all SM background coming from
the decay of heavy SM particles.
The Standard Model processes generally does not fulfill at the same time both constraints.
$Z$ pair production (by Drell-Yan or in the $t,u$ channels)
with $Z\rightarrow l^{+} + l^{-}$ give a higher multiplicity of leptons but with a
very low $E_{t}^{miss}$ (one of the leptons falling out the detector acceptance could mime a
$E_{t}^{miss}$ but will decrease the lepton multiplicity).
$W^{\pm}$ pair production (by Drell-Yan or in the $t,u$ channels) and the associated production of $W$'s with
$Z$'s give a higher $E_{t}^{miss}$ through the neutrinos but proportionally less leptons.
The $t \overline{t}$ production with $W^{\pm} \rightarrow \nu_l + l^{\pm}$ and the leptons arising from b jets
reconstructed as isolated gives a higher $E_{t}^{miss}$ but is generally supressed by the isolation criteria
of leptons. The associated production $Zjj$ is similar to the above mentioned cases.
\vskip -0.0cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./3lpt70all20sigma3.eps,width=12.0cm }}
\caption{\small Sensitive domain in the SUGRA parameter space.
For the explanation of the symbols see the text. The bricked domain is excluded by
theory and the cross hatched one is excluded by experimental measurements.}
\label{excl_reg}
\end{Fighere}
\vskip 0.5 cm
Fig.\ref{excl_reg} shows the domain of the sensitivity
one can obtain in 1 year of running with LHC at low luminosity.
Due to the limited number of simulated events (which is much less
than one can detect in one year) we have an uncertainty on this
region: full squares indicate the grid points where $S > 5$
at 99\% CL, no symbols at the grid points indicate
that $S < 5$ at 99\% CL, finally the open squares correspond to the
cases where one cannot make any of the two above statements.
\vskip 0.2cm
{\it \bf Energy scale of the SUGRA signal}
\vskip 0.2cm
If a SUSY signal manifests itself, its energy scale can be
determined in the case when $R$ parity is conserved from the
distribution of the quantity~\cite{SUGRAincl}:
\begin{equation}
M_{eff} = \sum_{i=1}^4 p_{t,i}^j + E_t^{miss}
\label{eq:meff}
\end{equation}
where $p_{t,i}^j$ are the 4 jets with the highest transverse momentum.
The missing transverse energy,
$E_t^{miss}$ originates mainly from the $\tilde\chi_1^0$ decay.
In the case of $R$ parity violation we should therefore
replace $E_t^{miss}$ by the decay products of the $\tilde\chi_1^0$,
i.e. we have used instead of the definition (\ref{eq:meff})
the following one:
\begin{equation}
M_{eff} = \sum_{i=1}^4 p_{t,i}^j + \sum_{i=1}^4 p_{t,i}^l + E_t^{miss}
\label{eq:meffR}
\end{equation}
where $p_{t,i}^l$ are the 4 leptons with the highest transverse momentum.
The distributions of $M_{eff}$ for the LHC point No 1 and 5
can be seen in Fig.\ref{meff_p1p5}. The point No 1 (3rd line in
Table \ref{tb:sigstat}) is associated with a high mass scale,
whereas the point No 5 (5th line in
Table \ref{tb:sigstat}) corresponds to a medium mass scale.
This is well reflected in Fig.\ref{meff_p1p5}.
\begin{Fighere}
\centering
\mbox{\epsfig{file=./meff_p1p5.eps,height=10cm,width=15cm }}
\caption{\small Distribution of $M_{eff}$ for the LHC points
No 1 and 5 after 3 years of LHC run at low luminosity.
The inclusive cuts used here for both points are ($\lambda_{122}=10^{-3}$):
$N_{leptons} \geq 4$ and $E_{t}^{miss} \geq 50 GeV$.
The maximum in these distributions ($M_{eff}^{max}$) depends strongly on the mass
parameters of the models being a good observable for the mass scale.}
\label{meff_p1p5}
\end{Fighere}
\vskip 1cm
In order to see how the $M_{eff}$
distributions are correlated with the SUGRA mass scale,
we have chosen randomly 30 points in the region
of $0 \le m_0 \le 500$ GeV and $0 \le m_{1/2} \le 500$ GeV.
At each points we have generated 10 000 events (see
2nd line in Table \ref{tb:sigstat}) and determined
the maximum of the $M_{eff}$
distributions: $M_{eff}^{max}$.
The events were selected requiring more than 3 leptons
and missing transverse energy higher than 50 GeV in an event.
In the first plot of Fig.\ref{meffmod}
we show the correlation of this quantity with the SUGRA mass scale,
$M_{SUGRA}$ which we have defined as being the highest mass of the strongly
interacting SUSY partners :
\begin{equation}
M_{SUGRA} = \mbox{min} (m_{\tilde g}, m_{\tilde q_R}, m_{\tilde b_1}, m_{\tilde t_1})
\label{eq:mSUGRA}
\end{equation}
The observed strong correlation is even more pronounced in the
second plot of Fig.\ref{meffmod}
where the distribution of the ratio of $M_{eff}^{max}/M_{SUGRA}$
is shown.
\vskip -1.cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./meff_models.eps,width=15cm }}
\vskip -7.0cm
\caption{\small Corellation between $M_{eff}^{max}$ and $M_{SUGRA}$ and
the distribution of the ratio $M_{eff}^{max}/M_{SUGRA}$ for 30 SUGRA points
(for other explanations see the text).}
\label{meffmod}
\end{Fighere}
\vskip 1cm
We have introduced another quantity to characterize the SUGRA energy scale
in the case of $R$ parity violation. Indeed, the algebraic sum of the
lepton transverse momentum divided by the number of leptons:
\begin{equation}
p_t^{l,norm} = (\sum_{i=1}^4 p_{t,i}^l)/N^l
\label{eq:ptnorm}
\end{equation}
gives a correlation with $m_0$ and $m_{1/2}$ as shown in Fig.\ref{ptnorm},
where the symbols delimite the regions where $p_t^{l,norm}$ is higher than
a certain value. On this plot we have selected events
with $N_{lept} \geq 3$,
$E_t^{miss} > 50$ GeV, and required a minimum value of 15 GeV for the
momentum of any of the leptons and 50 GeV for the leading one.
\vskip 0.2cm
{\it \bf Sensitivity for the type of the coupling}
\vskip 0.2cm
As mentioned in Section 2.1 counting the number of stable
leptons not only can reveal the signal of $R$ parity violation
but also can give a hint on the type of the coupling which
is realized. In order to show this we have simulated events
in three different points of the SUGRA space (LHC No 1,3 and 5,
see lines 3-5 of Table \ref{tb:sigstat}) and for two different
couplings: $\lambda_{122}$ and $\lambda_{123}$.
In the Table \ref{tb:4lept} the number of events are given
one can observe for the two classes: $0 e + 4 \mu$ and
$4 e + 0 \mu$ in one year of LHC running at low luminosity.
The events satisfy the following criteria: \\
\indent \indent $ 20 \le p_t^l \le 250$ GeV and $E_t^{miss} \ge 250$ GeV at points 1 and 5 \\
\indent \indent$ 10 \le p_t^l \le 100$ GeV and $E_t^{miss} \ge 150$ GeV at point 3.
\begin{Fighere}
\centering
\mbox{\epsfig{file=./ptnormlept2.eps,width=13.5cm }}
\caption{\small In the case of $\lambda_{123}=10^{-3}$, $A_0 = 0$, $tan \beta = 2$,
$sign \mu = +1$ and for the global cuts :
$N_{lept} \geq 3$, $E_{t}^{miss} \geq 50$ GeV, $p_{t}^{l} \geq 15$ GeV and
$p_{t}^{l, max} \geq 50$ GeV one can "divide" the $(m_{0}, m_{1/2})$ space in
$p_t^{l,norm}$ domains:
-- full squares: $p_t^{l,norm} \in (30, 150)$ GeV;
-- fat squares: $p_t^{l,norm} \in (150, 250)$ GeV;
-- open squares: $p_t^{l,norm} \in (250, 500)$ GeV. Bricked domain is excluded by theory and
the hatched one is excluded by the present experimental limits.}
\label{ptnorm}
\end{Fighere}
\begin{Tabhere}
\centering
\caption{\small Number of events with different number of electrons and muons }
\vskip 0.5cm
\begin{tabular}{|l|c|c|c|c|} \hline
& \multicolumn{2}{ c|}{ $\lambda_{122}$ } & \multicolumn{2}{ c|}{ $\lambda_{123}$ } \\ \cline{2-5}
& $0 e + 4 \mu$ & $4 e + 0 \mu$ & $0 e + 4 \mu$ & $4 e + 0 \mu$ \\ \hline
Point No 1 & $934 \pm 16$ & $3 \pm 1 $ & $76 \pm 4$ & $68 \pm 4$ \\ \hline
Point No 5 & $3010 \pm 52$ & $20 \pm 4 $ & $289 \pm 17$ & $304 \pm 18$ \\ \hline
Point No 3 & $52636 \pm 1760$ & $3356 \pm 444$ & $12400 \pm 600$ & $15565 \pm 672$ \\ \hline
\end{tabular}
\label{tb:4lept}
\end{Tabhere}
\vskip 0.5cm
It is obvious that one can clearly distinguish which of the two couplings
are realized in Nature. This is further illustrated in Fig.\ref{4lept},
where we have plotted the distribution of the number of muons/event for the two
different couplings where the total number of electrons and muons is
equal to four.\\
To conclude this section we have demonstrated that even if $R$ parity is violated
ATLAS can detect the signal of SUGRA in a large domain of the parameter space.
This domain is compatible with that obtained in the case when $R$ parity is
conserved. We can equally well establish the mass scale of the SUGRA compared
to the case of $R$ parity conservation. Finally, one has a possibility to
determine the type of the coupling if only one causes the violation of the
$R$ parity.
\begin{Fighere}
\centering
\mbox{\epsfig{file=./multipleff2.eps,width=15cm }}
\vskip -0.5 cm
\caption{\small The distribution of the number of muons per event if the total number
of electrons and muons is equal to 4 in the case of two $\rlap/R$ couplings ($\lambda_{122}$,
$\lambda_{123}$) in the SUGRA points 1, 3 and 5.}
\label{4lept}
\end{Fighere}
\vspace {0.5cm}
\newpage
\section{Exclusive measurements}
\vskip 0.5 cm
In this chapter we show that one can determine the parameter
values of the SUGRA model, similarly to the case when $R$ parity is
conserved~\cite{SUGRA12}-\cite{SUGRA5}.
The quantities to be used for this purpose, in general,
are the masses of the reconstructed SUSY particles as well
as their observed production cross sections and branching ratios.
We shall use only the first type of characteristics in this
analysis since determination of cross sections and branching ratios
is more sensitive to the acceptances and in many cases detailed
simulations are needed. In 3 out of the 5 LHC points, which are sufficiently different
to illustrate the methods in various conditions, we shall first
show how to reconstruct the SUSY particles, determine their masses,
and finally fit the model parameters to these mass values.
The reconstruction of the SUSY particles are difficult because there
are always at least two final state particles which cannot be detected.
These are obviousely the LSP's in the case of $R$ parity conservation,
but if $R$ parity is violated through the terms (\ref{eq:Lagl}) there is
always at least one undetectable neutrino produced
in each LSP decay (see also Table \ref{tb:BRs}).
To overcome this difficulty we use the fact that in a 3-body decay:
\begin{equation}
A \longrightarrow a\ b\ c
\label{eq:Aabc}
\end{equation}
the invariant mass $m_{bc}$ of two of the three final state particles, e.g.
$b$ and $c$ gives a clear endpoint. The endpoint, $m_{bc}^{end}$
is related to the masses of $A$ and that of the undetected particle, $a$:
\begin{equation}
m_{bc}^{end} = m_A - m_a
\label{eq:mAa}
\end{equation}
since in the restframe of $A$ the three-momentum of $a$ (or that
of the ($bc$) system) is zero.
This equation is particularly useful if the mass of $a$ is zero: in this
case one can directly estimate the mass of the particle $A$ by measuring
the endpoint of the $m_{bc}$ distribution. Of course, in the practice
this measurement is difficult because the endpoint may be hidden
by eventual background and it can disappear if either $b$ or $c$
is unstable. This is the case when a $\tau$ is produced in the decay
of the $\tilde\chi_1^0$, thus in all couplings when a "3" appears in
the index of the $\lambda$.
Once the endpoint is established one can determine even the four-momentum
$P^{\mu}$ of the undetected particle $a$ or that of the mother particle $A$
by selecting events around the endpoint:
\begin{equation}
P^{\mu}_{A,a} = \frac{m_{A,a}}{m_{bc}}(P^{\mu}_{b}+P^{\mu}_{c}) \ (\mu = 1,...,4)
\label{eq:pAa}
\end{equation}
As one can see from Equs. (\ref{eq:mAa}) and (\ref{eq:pAa}),
if particle $a$ has zero mass (as e.g. in the case of a neutrino),
the four-momentum of the charged lepton pair is equal to that of particle $A$.
The four-momentum of particle $A$ can be used further for the reconstruction
of the parent particle, and so on, until one arrives at the beginning of the
decay chain, i.e. at the original SUSY particle. It is also obvious that at the endpoint
the 4-momentum of particle $a$ and $A$ are parallel whatever their masses are.
Since equations (\ref{eq:mAa}) and (\ref{eq:pAa}) are strictly valid
only at the endpoint, one usually needs a large number of produced events
and in the selection of the size of the region around the endpoint
an optimum has to be found between the statistical error and
the approximate validity of the above relations. In practice one applies in
Eq.(~\ref{eq:pAa}) $m_{bc}$ instead of $m^{end}_{bc}$ in order to "rescale" the
values of $P^{\mu}_b+P^{\mu}_c$ which are smaller than their corresponding
values at the endpoint.
If the decay (\ref{eq:Aabc}) proceeds through a sequence of 2 two-body decays:
\begin{equation}
A \longrightarrow B\ b \ \ \ \ \ \ \ \ B \longrightarrow a\ c
\label{eq:AaBbc}
\end{equation}
one can observe an endpoint in the $m_{bc}$ distribution, whose value is given by:\\
\begin{equation}
m^{end}_{bc}=m_A \sqrt{1-(\frac{m_B}{m_A})^2}\ \sqrt{1-(\frac{m_a}{m_B})^2}\mbox{\ \ .}
\label{eq:mend22b}
\end{equation}
\\
The quantity of $\sigma \times BR$ of the produced particle $A$ helps to
disentangle which of the two decay modes (\ref{eq:Aabc}) or (\ref{eq:AaBbc}) has
occured and to choose between (\ref{eq:mAa}) or (\ref{eq:mend22b}) to estimate the masses.
At LHC the generic production and decay chains are represented in the schemas here below: (*).
The squark or gluino undergoes the cascade decay:
\begin{tabbing}
oooooooooooooooooooooo\=oooooooooooo \kill
\> $\tilde q\longrightarrow$ \= $\tilde g + q$ \\
\> \>
\begin{picture}(11,10)(0,0)
\put(2.0,9.0){\line(0,-2){8}}
\put(2.0,1.0){\vector(1,0){9}}
\end{picture}
\= $\tilde q'$ + $q'$ \\
\> \>\>
\begin{picture}(11,10)(0,0)
\put(2.0,9.0){\line(0,-2){8}}
\put(2.0,1.0){\vector(1,0){9}}
\end{picture}
\= $\tilde \chi_1^{\pm}$ + $q''$ \\
\> \>\>\>
\begin{picture}(11,10)(0,0)
\put(2.0,9.0){\line(0,-2){8}}
\put(2.0,1.0){\vector(1,0){9}}
\end{picture}
\=$\tilde \chi_1^{0}$ + $W^{\pm}$ oooooooooooooooooooo\= (*) \kill
\> \>\>\>
\begin{picture}(11,10)(0,0)
\put(2.0,9.0){\line(0,-2){8}}
\put(2.0,1.0){\vector(1,0){9}}
\end{picture}
\=$\tilde \chi_1^{0}$ + $W^{\pm}$ \> (*) \\
\> \>\>\>\>
\begin{picture}(11,10)(0,0)
\put(2.0,9.0){\line(0,-2){8}}
\put(2.0,1.0){\vector(1,0){9}}
\end{picture}
$l^+$ + $l^-$ + $\nu$ \\
\> \>\>
\begin{picture}(11,10)(0,0)
\put(2.0,9.0){\line(0,-2){8}}
\put(2.0,1.0){\vector(1,0){9}}
\end{picture}
\= $\tilde \chi_2^0$ + $q$ \\
\> \>\>\>
\begin{picture}(11,10)(0,0)
\put(2.0,9.0){\line(0,-2){8}}
\put(2.0,1.0){\vector(1,0){9}}
\end{picture}
$\tilde \chi_1^{0}$ + $(l^+l^-)/h^0$ \\
\> \>\>\>\>
\begin{picture}(11,10)(0,0)
\put(2.0,9.0){\line(0,-2){8}}
\put(2.0,1.0){\vector(1,0){9}}
\end{picture}
$l^+$ + $l^-$ + $\nu$ \\
\end{tabbing}
and the $R$ parity violation occurs at the end of the decay chain in the
decay of the $\tilde \chi_1^{0}$.
For the exclusive measurements we have chosen 3 out of the 5 LHC points to
study. Point 1 has a high, point 3 has a low and point 5 has a medium mass scale.
Another crucial difference between these points, coming from
the mass parameters $m_0$, $m_{1/2}$ and sign($\mu$), is reflected
in the decay of $\tilde\chi_2^0$.
In point 3, $\tilde\chi_2^0$ decays mainly through a three body decay
$\tilde\chi_2^{0} \rightarrow \tilde\chi_1^0 + l^{\pm} + l^{\mp}$
(with a virtual $Z$ or $\tilde l$).
In points 1 and 5 the main decay ($65 \% \div 80 \%$) of $\tilde\chi_2^0$ proceeds
through a two-body decay $\tilde\chi_2^{0} \rightarrow \tilde\chi_1^0 + h^{0}$,
but also with a non-negligible branching ratio ($20 \% \div 30 \%$)
it decays in a sequence of 2 two-body decays
$\tilde\chi_2^{0} \rightarrow \tilde \l_{R}^{\pm} + l^{\mp} \rightarrow
\tilde\chi_1^0 + l^{\pm} + l^{\mp}$ with a $\tilde l_R^{\pm}$ being real.
Therefore point 3 has a completely different event topology,
namely a very large number of leptons in the final
state, in comparison with the other two points. This is
the reason that we present here an analysis in point 3 although
it is already excluded by the LEP limit on the Higgs mass~\cite{LEP_Higgs}.
In a nearby point, which is not yet excluded all our conclusions can
be considered as valid.
At each LHC points we have considered two different couplings:
$\lambda_{122}$ and $\lambda_{123}$. In the case of the
first one the reconstruction of the SUSY particles is easier
than in the second one, where a $\tau$ particle appears
always among the decay products. This means more missing energy
in the second case and also the absence of a clear endpoint
in the invariant mass of the opposite-sign and different flavor (OSDF) lepton pair.
This in turn makes very difficult if not
impossible the reconstruction of the $\tilde\chi_1^0$ .
The values of the SUGRA parameters as well as the number
of generated events is listed in Table\ref{tb:sigstat}.
The masses of the SUSY particles in the three points are
listed in Table~\ref{tb:masses}.
\vskip 0.5 cm
\subsection{The LHC points No1 and No5}
\vskip 1.0 cm
Concerning the reconstruction of the SUSY particles these two points
are very similar and therefore we treat them here together.
In the cascade decay chain (*) the predominant decay of the
$\tilde\chi_2^0$ proceeds via
$\tilde\chi_2^0\longrightarrow h^0 + \tilde\chi_1^0$
with a BR of $\sim 99 \%$ in point 1 and $\sim 63 \%$ in point 5,
where the competitive decay through $\tilde l_{R}$ becomes also important.
The main inclusive cuts to select SUSY events are based on the requirement of a high
lepton multiplicity, $N_l \ge 4$ and a moderate missing transverse energy,
$E_t^{miss}$.
The SM background is small and can be completely neglected
applying a cut on $E_t^{miss}$
as is shown in Fig.\ref{etm_p15}
where the $E_t^{miss}$ distribution is shown for the points 1
and 5 and for $\lambda_{122}$ and $\lambda_{123}$.
\vspace{-0.5cm}
\begin{Tabhere}
\centering
\caption{\small SUSY particle masses at points 1, 3 and 5 in GeV}
\vskip 0.5cm
\begin{tabular}{|l|c|c|c|} \hline
LHC point & 1 & 3 & 5 \\ \hline
$\tilde g$ & 1008 & 299 & 769 \\ \hline
$\tilde q_L$ & 958 & 317 & 687 \\ \hline
$\tilde q_R$ & 925 & 312 & 664 \\ \hline
$\tilde b_2$ & 922 & 313 & 662 \\ \hline
$\tilde b_1$ & 855 & 278 & 634 \\ \hline
$\tilde t_2$ & 913 & 325 & 706 \\ \hline
$\tilde t_1$ & 649 & 260 & 494 \\ \hline
$\tilde l_L$ & 490 & 216 & 239 \\ \hline
$\tilde l_R$ & 430 & 207 & 157 \\ \hline
$\tilde \nu_L$ & 486 & 207 & 230 \\ \hline
$\tilde \tau_2$ & 490 & 216 & 239 \\ \hline
$\tilde \tau_1$ & 430 & 206 & 157 \\ \hline
$\tilde \chi^{\pm}_2$ & 775 & 274 & 526 \\ \hline
$\tilde \chi^{\pm}_1$ & 326 & 96 & 232 \\ \hline
$\tilde \chi^0_4$ & 778 & 275 & 529 \\ \hline
$\tilde \chi^0_3$ & 762 & 258 & 505 \\ \hline
$\tilde \chi^0_2$ & 326 & 97 & 233 \\ \hline
$\tilde \chi^0_1$ = LSP & 168 & 45 & 122 \\ \hline
$h_0$ & 98 & 69 & 95 \\ \hline
\end{tabular}
\label{tb:masses}
\end{Tabhere}
\vskip 0.5 cm
The "structures" seen in the hatched histograms of Fig.\ref{etm_p15} is due to
statistical fluctuations.
In what follows we shall demonstrate that in the case of the
coupling $\lambda_{122}$ we can reconstruct all the SUSY particles
in the decay chain (*), starting with the $\tilde\chi_1^0$,
and in this respect we can achieve more than it was the case
with $R$ parity conserved. On the other hand, if $\lambda_{123}$
is nonzero, we can reconstruct only a fraction of the SUSY particles,
and the decay products of the $\tilde\chi_1^0$, i.e. the additional
leptons, increase the background and thereby deteriorate the determination
of some particle masses. We shall show that in spite of these
difficulties the achieved precision is at worst comparable with
the one in the case of conserved $R$ parity.
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p1p5etmis122.eps,width=7.cm } \epsfig{file=./p1p5etmis123.eps,width=7.cm }}
\caption{\small The distribution of $E_t^{miss}$ for SUSY events (open histogram) and SM events
(hatched histogram) in points 1 and 5 for $\lambda_{122}=10^{-3}$ (left plots) and
$\lambda_{123}=10^{-3}$ (right plots) for events with $N_{leptons} \geq 4$, $P_{t}^{lept} \geq 10$ GeV
and after 3 years of LHC run at low luminosity. The arrows precise the cuts to be used.}
\label{etm_p15}
\end{Fighere}
\vskip 0.8cm
{\bf \large \it \underline{The case $\lambda_{122} \ne 0$}}
\vskip 0.2cm
{\it \bf Reconstruction of $\tilde \chi^0_1 \rightarrow \nu_{e (\mu)} + e^{\pm} (\mu^{\pm}) + \mu^{\mp} $}
\vskip 0.2cm
For the reconstruction of the $\tilde\chi_1^0$ we apply
the following selection criteria: \\
\indent \indent $(i)$ $N_l \ge 4$ (and $N_l^+ = N_l^-$ and $N_e = N_{\mu}$ for point 5), \\
\indent \indent $(ii)$ $p_t^l \ge $ 10 GeV, cos($\alpha_{l^{\pm}l^{'\mp}}$) $\geq$ 0.5, \\
\indent \indent $(iii)$ $E_t^{miss} \ge $ 50 GeV , \\
where $\alpha_{l^{\pm}l^{'\mp}}$ is the angle between any OSDF (i.e. electron -muon) leptons.\\
The invariant mass distributions of the OSDF lepton pairs
are shown in Fig.\ref{p1chi01} for point 1
and in Fig.\ref{p5chi01} for point 5, where the number of events correspond to
3 years of LHC run at low luminosity. \\
In point 1, isolated leptons are produced practically only in the two $\rlap/R$ decays of the
$\tilde \chi^0_1$'s, therefore the background of OSDF distribution will be mainly
$\tilde \chi^0_1$ combinatorial.
In point 5, as we mentioned before, a non-negligible production of isolated leptons (beyond
the $\rlap/R$ decays of $\tilde \chi^0_1$'s) comes from
$\tilde \chi^{0}_{2} \rightarrow \tilde l^{\pm}_{R} + l^{\mp} \rightarrow
\tilde \chi^{0}_{1} + l^{\pm} + l^{\mp}$. This decay will produce two leptons of opposite sign (OS)
but of the same flavor (SF), changing the balance of leptons per flavor.
Most of the events with an increased number of leptons, will have only one $\tilde \chi^{0}_{2}$
decaying leptonically (the other one, if it exists, will decay in $h^{0}$'s). To decrease the
number of bad combinations (with leptons not coming from $\tilde \chi^{0}_{1}$'s), additional cuts
like $N_l^+ = N_l^-$ and $N_e = N_{\mu}$ in each event are very efficient.
There is a clear endpoint over a moderate
background at both points 1 and 5 corresponding to the
$\tilde\chi_1^0$ mass (c.f. Table\ref{tb:masses}) in virtue
of Equ.(\ref{eq:mAa}).
We parametrize and subtract the background with a
Maxwellian distribution and fit the resulting distribution near the
endpoint obtaining the $\tilde\chi_1^0$ mass values:
\begin{equation}
m_{\tilde\chi_1^0}^{meas} = 169.80^{+0.2}_{-0.8} \mbox{\ GeV at point 1}
\label{eq:mchi10p1}
\end{equation}
\begin{equation}
m_{\tilde\chi_1^0}^{meas} = 122.62^{+0.4}_{-1.0} \mbox{\ GeV at point 5}
\label{eq:mchi10p5}
\end{equation}
\vskip -3.0 cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p1mass_chi01.eps,height=9.0cm,width=15cm }}
\vskip -0.5 cm
\caption{\small The invariant mass distribution of the OSDF
lepton pairs for 3 years of LHC run at low luminosity at point 1 (for the selection
criteria see text). In the upper-left plot are represented all events (open histogram)
and the background (shadded histogram) which is mainly SUSY combinatorial.
The upper-right and lower-left plots show a Maxwellian-like distribution fitted to the background.
The lower-right plot represents the result after the subtraction of the fitted background.
The edge is fitted with a polinomial function and the error is only statistical. }
\label{p1chi01}
\end{Fighere}
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p5mass_chi01.eps,height=9.0cm,width=15cm }}
\vskip -0.5 cm
\caption{\small The invariant mass distribution of the OSDF
lepton pairs for 3 years of LHC run at low luminosity at point 5 (for the selection
criteria see text). }
\label{p5chi01}
\end{Fighere}
\vskip 1.0 cm
The estimated error contains both statistical and systematical.
This latter is due to the energy resolution of the leptons
and mainly to the finite bin-size of our histograms around the endpoint.
Since due to the bin-size one tends to overestimate the mass value the systematical
error is asymetrical.\\
For any further reconstruction using OSDF pairs (like $\tilde \chi^0_2$, $\tilde \chi^{\pm}_1$, etc)
one will take as $\tilde \chi^0_1$ candidates the OSDF pairs with an invariant mass in
($m_{\tilde \chi^0_1}-\Delta m_{\tilde \chi^0_1}$, $m_{\tilde \chi^0_1}$), where
$\Delta m_{\tilde \chi^0_1}=50$ GeV for point 1 and $\Delta m_{\tilde \chi^0_1}=30$ GeV for point 5.
To improve the statistics in point 5 we will not use the additional cuts (see $(i)$).
\vskip 0.2cm
{\it \bf Reconstruction of $h^0 \rightarrow b + \bar b$}
\vskip 0.2cm
To the global cuts ($N_l \geq 4$, $P_{t}^{l} \geq 10$ GeV and $E_t^{miss} \geq 50$ GeV)
we will add some specific cuts on b jets.
Since the $h^0$ decays to $b\bar b$ pairs with appr. 88\% BR its reconstruction
proceeds by the selection of these pairs, adding thus the following detection
criteria to the above ones: \\
\indent \indent $(iv)$ $p_t^b \ge 30$ GeV for point 1, $40$ GeV for point 5 respectively,
and $p_t^b \le $ 300 GeV for both points \\
\indent \indent $(v)$ cos($\alpha_{b\bar b}$) $\ge$ 0.4 (point 1) or 0.3 (point 5) \\
where $\alpha_{b\bar b}$ is the angle between the $b$ and $\bar b$.
The obtained invariant mass distributions of the $b\bar b$ pairs are shown
in Fig.\ref{p1h0mass} and Fig.\ref{p5h0mass}. There is a clear mass peak corresponding
to the $h^0$ particle (c.f. Table\ref{tb:masses}).
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p1p5bsource.eps,width=12cm,height=7cm }}
\vskip -0.5 cm
\caption{\small The sparticle mothers of b jets (in Pythia codes) in the points 1 and 5
(see text).}
\label{p1p5bsource}
\end{Fighere}
\vskip 0.5 cm
As we can see from the Fig.\ref{p1p5bsource}, the b jets are mainly produced through
$h^0 \rightarrow b \bar b$ but also through
$\tilde g \rightarrow \tilde b_{1(2)} + \bar b \rightarrow \tilde \chi^0_{1(2)} + b + \bar b$ or
$\tilde g \rightarrow \tilde t_1 + \bar t \rightarrow \tilde \chi^0_{1(2)} + t + \bar t$
(or $\rightarrow \tilde \chi^{+}_{1} + b + \bar t$) and the subsequent decay of $t$.
In the case of conserved $R$ this SUSY background (as well as the SM $t \bar t$) is efficiently
rejected with a veto on leptons (coming from the top decay). However, in the case of
$\rlap/R$ through $\rlap/L$ couplings we cannot apply anymore such a cut (and a veto on
{\it additional} leptons doesn't increase significantly the $S/B$ ratio).
Another source of background (albeit small) is the production
of $Z$'s (i.e. $\tilde t_{2} \rightarrow \tilde t_{1} + Z^0$ or
$\tilde \chi^{0}_{i} \rightarrow \tilde \chi^{0}_{1} + Z^{0}$, $i=2,3,4$
or $\tilde \chi^{\pm}_{2} \rightarrow \tilde \chi^{\pm}_{1} + Z^0$, with $Z^0 \rightarrow b + \bar b$)
resulting in a peak very close to the $h^0$ one.
This will give rise to an asymmetric $h^0$ peak with a higher width as can be seen in both
points. A similar effect arises also from the calibration error of $b$ jets.
After havig parametrized and subtracted this
background we have fitted the mass peak with a Gaussian curve
and obtained the values: \\
\begin{equation}
m_{h^0}^{meas} = 97.08 \pm 1.5 \mbox{\ GeV at point 1}
\label{eq:mh0p1}
\end{equation}
\begin{equation}
m_{h^0}^{meas} = 94.7 \pm 1.5 \mbox{\ GeV at point 5}
\label{eq:mh0p5}
\end{equation}
The estimated systematic error is mainly due to the uncertainties in the
$b$-jet correction factor shown in Fig.\ref{p5selbjets}.
\vskip -0.5 cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p1_h0mass.eps,height=10.0cm,width=15cm }}
\vskip -5. cm
\caption{\small The invariant mass distribution of the $b\bar b$
pairs for 3 years of LHC run at low luminosity at point 1.
On the left plot are represented all the events and the background (shadowed).
The background is fitted with a Maxwellian function and subtracted. The result
is drawn in the right plot. The arrow points to the theoretical value of $m_{h^0}$.
The peak is fitted with a gaussian. The written error contains only the statistical part.}
\label{p1h0mass}
\end{Fighere}
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p5_h0mass.eps,height=10.0cm,width=15cm }}
\vskip -5. cm
\caption{\small The invariant mass distribution of the $b\bar b$
pairs for 3 years of LHC run at low luminosity at point 5.}
\label{p5h0mass}
\end{Fighere}
\vskip 1cm
\vskip 0.2cm
{\it \bf Reconstruction of $\tilde \chi^0_2 \rightarrow h^{0} + \tilde \chi^0_1$}
\vskip 0.2cm
The reconstructed $h^0$ and $\tilde\chi_1^0$ allows the reconstruction of
the mother particle, the $\tilde\chi_2^0$, since this latter decays to
the formers by 99\% BR at point 1 and 63\% BR at point 5.
We will require the $\tilde \chi^0_1$ candidates (OSDF lepton pairs) to be "close" to the endpoint :\\
\indent \indent $(vi)$ $m(OSDF) \in (m^{endp}-50,m^{endp})$ GeV - in point 1 and,\\
\indent \indent \ \ \ \ \ \ $m(OSDF) \in (m^{endp}-30,m^{endp})$ GeV - in point 5,\\
and the $h^0$ candidates ($bb$ pairs) to be around the mass peak of $h^0$ :\\
\indent \indent $(vii)$ $m(bb) \in (m^{peak}-15,m^{peak}+15)$ GeV - for both points 1 and 5.\\
Looking at the relations between $\tilde \chi^0_2$, $\tilde \chi^0_1$ and $h^0$ masses, one can expect
in point 1 a higher boost of $\tilde \chi^0_1 - h^0$ pair, comparing to point 5, therefore we demand :\\
\indent \indent $(viii)$ cos($\alpha_{h_0\tilde\chi_1^0}$) $\ge$ 0.7 (point 1) or 0.5 (point 5). \\
Finally one gets the invariant mass distributions of the $h_0\tilde\chi_1^0$ pairs as
shown in the Fig.\ref{mh0chi10}.
The background in this case is combinatorial on the one hand, and comes from SUSY production
of $t \bar t$ pairs with :
$t \bar t \rightarrow b + W^{+} + \bar b + W^{-} \rightarrow b + \bar b + l^{+} + l^{, -} +
\nu_l + \bar \nu_{l^{,}} $ on the other hand.
However, the strong correlation required separately between the leptons and between the b jets
will supress it considerably.
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p1p5_chi02.eps,width=14cm }}
\vskip -7.0 cm
\caption{\small The invariant mass distribution of the $h_0\tilde\chi_1^0$
pairs for 3 years of LHC run at low luminosity in points 1 and 5.
The tail is due to bad combinations with the wrong $\tilde \chi^0_1$.
The arrows point to the theoretical values of $m_{\tilde \chi^0_2}$.
The masses are determined by a gaussian fit around the peak.
The error is only the statistical.}
\label{mh0chi10}
\end{Fighere}
\vskip 1.0 cm
The peak around the $\tilde\chi_2^0$
mass value (c.f. Table~\ref{tb:masses}) allows to estimate this latter
by fitting a Gaussian curve on the peak:
\begin{equation}
m_{\tilde\chi_2^0}^{meas} = 326.2 \pm 6 \mbox{\ GeV at point 1}
\label{eq:mchi20p1}
\end{equation}
\begin{equation}
m_{\tilde\chi_2^0}^{meas} = 230.7 \pm 3.9 \mbox{\ GeV at point 5}
\label{eq:mchi20p5}
\end{equation}
\vskip 1.5cm
{\it \bf Reconstruction of $\tilde \chi^{\pm}_1 \rightarrow \tilde \chi^0_1 + W^{\pm}$}
\vskip 0.2cm
The same technique allows the reconstruction of the lightest chargino ($\tilde \chi^{\pm}_1$).
As one can see in the scheme (*) one reconstructs
first the $W^{\pm}$.
The $W$ reconstruction is carried out by combining light quark jet pairs
using the additional selection criteria (to ($i \div iii$) and ($vi$)): \\
\indent \indent $(ix)$ $p_t^j \ge $ 100 GeV and $p_t^j \le $ 600 GeV (point 1) or 350 GeV (point 5) \\
\indent \indent $(x)$ cos($\alpha_{jj}$) $\ge$ 0.9 (point 1) or 0.87 (point 5) \\
where $\alpha_{jj}$ is the angle between the $jj$ pair.
The obtained invariant mass distributions of the $jj$ pairs are shown
in Fig.\ref{mcha1}. There is a clear mass peak at the place of the $W$ mass.
Selecting the $W$ candidates around the $W$ mass peak ($\pm 15$ GeV)
and using the $\tilde\chi_1^0$ candidates close to the
endpoint (see $(vi)$) we plot
the invariant mass distribution of the $\tilde\chi_1^0 W$ pairs
in Fig.\ref{mcha1}, by requiring that the $\tilde\chi_1^0$ and the $ W$
be close in phase space: \\
\indent \indent $(xi)$ cos($\alpha_{W\tilde\chi_1^0}$) $\ge$ 0.85 in both points \\
where $\alpha_{W\tilde\chi_1^0}$ is the angle between the $\tilde\chi_1^0 W$ pair.
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p1p5_cha1.eps,width=14.cm }}
\caption{\small The invariant mass distribution of the $jj$
pairs for 3 years of LHC run at low luminosity at point 1
and 5 (upper plots). Selecting W candidates arround $W$ mass peak ($\pm 15$ GeV
- between the bars), we combine them with the $\tilde \chi^0_1$ candidates and obtain the
distributions represented in the lower plots for point 1 and 5 respectively.
The arrows point to the theoretical values for $m_{\tilde \chi^{\pm}_1}$.}
\label{mcha1}
\end{Fighere}
\vskip 1.0 cm
One can see a clear peak at the mass value of the lightest chargino at point 5
(c.f. Table\ref{tb:masses}). At point 1, where the statistics is much more
limited due to the heaviness of the chargino, the corresponding peak
is less clear. We have determined the mass of the chargino by fitting a
Gaussian around the mass peaks which gives:
\begin{equation}
m_{\tilde\chi_1^{\pm}}^{meas} = 328.2 \pm 6.5 \mbox{\ GeV at point 1}
\label{eq:mchap1}
\end{equation}
\begin{equation}
m_{\tilde\chi_1^{\pm}}^{meas} = 232.2 \pm 4.5 \mbox{\ GeV at point 5.}
\label{eq:mchap5}
\end{equation}
\vskip 1.5cm
{\it \bf Reconstruction of $\tilde q_{R} \rightarrow \tilde \chi^{0}_{1} + jet$ }
\vskip 0.2cm
A competitive cascade corresponding to this decay in (*) is :\\
\begin{tabbing}
oooooooooooooooooooooo\=oooooooooooo \kill
\>( \ \ $\tilde g \longrightarrow$ \= $\tilde q_R + q$ \ \ ) \\
\> \>
\begin{picture}(11,10)(0,0)
\put(2.0,9.0){\line(0,-2){8}}
\put(2.0,1.0){\vector(1,0){9}}
\end{picture}
\= $\tilde \chi^0_1 + q$ \indent \indent \indent \indent \indent (* *) \\
\> \>\>
\begin{picture}(11,10)(0,0)
\put(2.0,9.0){\line(0,-2){8}}
\put(2.0,1.0){\vector(1,0){9}}
\end{picture}
$l^{\pm}$ + $l^{\mp}$ + $\nu$ \\
\end{tabbing}
Due to the fact that the squarks are very heavy, the jet produced
in the $\tilde q_R$ decay will carry an important fraction of energy.
To avoid high combinatorial background one can ask only for one very energetic
light jet in the event.
Applying the following additional selection criteria (to ($i \div iii$) and ($vi$))
for the light quark jets: \\
\indent \indent $(xii)$ One light jet with $p_t^j \ge 750$ GeV (point 1) or 400 GeV (point 5) whose\\
\indent \indent \ \ \ \ \ \ \ invariant mass with any other light jet is outside the\\
\indent \indent \ \ \ \ \ \ \ $W^{\pm}$ or $Z^0$ masses ($\pm 15$ GeV);\\
\indent \indent $(xiii)$ cos($\alpha_{j\tilde\chi_1^0}$) $\ge 0.0$ in both points \\
where $\alpha_{j\tilde\chi_1^0}$ is the angle between the $j \tilde\chi_1^0$ pair.
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p1p5_sqr.eps,width=14cm }}
\vskip -0.5 cm
\caption{\small The invariant mass distribution of the $j\tilde\chi_1^0$
pairs for 3 years of LHC run at low luminosity at point 1
and at point 5. The vertical arrows point to the nominal values of masses for different sparticles.}
\label{p1p5_sqr}
\end{Fighere}
\vskip 0.5 cm
The obtained invariant mass distributions of the $j\tilde\chi_1^0$ pairs are shown
in Fig.\ref{p1p5_sqr}.
The $\tilde q_R$ can be produced directly in the hard process or in the
decay of $\tilde g$. In points 1 and 5 the jet associated with $\tilde q_R$ is very soft
comparing to the jet associated with $\tilde \chi^0_1$ and often the soft jet can fall in the
cone of the reconstructed hard jet. This is the reason of the presence of a shoulder at the
$\tilde g$ mass value in the distributions of Fig.\ref{p1p5_sqr}.
Another source of background is the decay
$\tilde q_L \rightarrow \tilde \chi^{\pm}_1 + q$ with
$\tilde \chi^{\pm}_1 \rightarrow \tilde \chi^0_1 + W^{\pm} \rightarrow \tilde \chi^0_1 + \nu_l + l^{\pm}$
and $l^{\pm}$ undetected.
The left shoulder at the $\tilde t_1$ mass at point 5 corresponds to the case where the 3 jets
of the $\tilde t_1 \rightarrow \tilde \chi^{+}_1 + b \rightarrow \tilde \chi^0_1 + j + j' + b$
or $\tilde t_1 \rightarrow \tilde \chi^{0}_1 + t \rightarrow \tilde \chi^0_1 + j + j' + b$
decay chains are erroneosely combined into a single energetic one.
One can see mass peaks at the values of the right handed
squarks (c.f. Table\ref{tb:masses}). The obtained mass values will
have large errors due to the superposition of all these effects :
\begin{equation}
m_{\tilde q_R}^{meas} = 932 \pm 20 \mbox{\ GeV at point 1}
\label{eq:msqrp1}
\end{equation}
\begin{equation}
m_{\tilde q_R}^{meas} = 662 \pm 12 \mbox{\ GeV at point 5}
\label{eq:msqrp5}
\end{equation}
\vskip 0.5cm
{\it \bf Reconstruction of $\tilde q_{L} \rightarrow \tilde \chi^{0}_{2} + jet$ in point 5}
\vskip 0.2cm
For the point 5 we can do more due to the higher statistics.
As we already mentioned in the case of $\tilde q_R$, the $\tilde q_L$ can decay with
a branching fraction of about 31 \% in $\tilde \chi^0_2$ and a light jet. This jet
is also very energetic.
In addition to the cuts $(i \div viii)$ one will select $\tilde \chi^0_2$ candidates from the
$\tilde \chi^0_1 h^0$ distribution :\\
\indent \indent $(xiv)$ $m(\tilde \chi^0_1 h^0) \in (m_{\tilde \chi^0_2}-40,m_{\tilde \chi^0_2}+40)$ GeV \\
and only one hard light-flavoured jet with :\\
\indent \indent $(xv)$ $p_t^{jet} \geq 100$ GeV whose invariant mass with any other
light jet is $\pm 15$ GeV \\
\indent \indent \ \ \ \ \ \ outside of the $W^{\pm}$ or $Z^0$ masses.\\
\vskip -0.5 cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p5_sql.eps,width=9.0 cm }}
\caption{\small The invariant mass distribution of the $j\tilde\chi_2^0$
pairs for 3 years of LHC run at low luminosity at point 5. Vertical arrows point to
the nominal values of $\tilde q_L$ and $\tilde t_1$ masses. For cuts see text.}
\label{p5_sql}
\end{Fighere}
The decay $\tilde t_1 \rightarrow \tilde \chi^0_2 + t \rightarrow \tilde \chi^0_2 + j + j + b$
when the 3 jets are erroneusely combined into a single hard jet, manifests itself as a peak
at the $\tilde t_1$ mass (see Fig.\ref{p5_sql}).
The gaussian fits to these peaks, after three years of LHC run at low luminosity, result in :\\
\begin{equation}
m_{\tilde q_L}^{meas} = 685 \pm 20 \mbox{\ GeV at point 5}
\label{eq:msqlp5}
\end{equation}
\begin{equation}
m_{\tilde t_1}^{meas} = 504 \pm 20 \mbox{\ GeV at point 5.}
\label{eq:mst1p5}
\end{equation}
\vskip 0.5cm
{\it \bf Reconstruction of $\tilde l^{\pm}_{R} \rightarrow \tilde \chi^{0}_{1} + l^{\pm}$ in point 5}
\vskip 0.2cm
In the point 5, the $\tilde l^{\pm}_{R}$ decays $\sim$ 100 \% in
$\tilde \chi^{0}_{1}$ and $l^{\pm}$.
In order to obtain the mass of the right handed slepton, we have combined
the $\tilde\chi_1^0$ candidates with a "soft" lepton not participating in the
$\tilde\chi_1^0$ reconstruction and satisfying: \\
\indent \indent $(xvi)$ $ 10 \le p_t^l \le $ 200 GeV \\
\indent \indent $(xvii)$ cos($\alpha_{l\tilde\chi_1^0}$) $\ge$ 0.5 \\
where $\alpha_{l \tilde\chi_1^0}$ is the angle between the $l \tilde\chi_1^0$ pair.
The obtained invariant mass distribution is shown in Fig.\ref{p5_slr}.
Besides the combinatorial one, the background arises from the decay
$\tilde \chi^{\pm}_1 \rightarrow \tilde \chi^0_1 + W^{\pm} \rightarrow \tilde \chi^0_1 + l^{\pm} + \nu_l$,
which results in an {\it endpoint}
in the $l^{\pm} \tilde \chi^0_1$ distribution at the mass of $\tilde \chi^{\pm}_1$.
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p5_slr.eps,width=9.5cm }}
\caption{\small The invariant mass distribution of the $l\tilde\chi_1^0$
pairs for 3 years of LHC run at low luminosity at point 5.
The arrows point to the nominal values of $m_{\tilde l_R}$ and $m_{\tilde \chi^{\pm}_1}$. }
\label{p5_slr}
\end{Fighere}
A gaussian fit to the mass peak gives :
\begin{equation}
m_{\tilde l_R}^{meas} = 156.8 \pm 1.8 \mbox{\ GeV at point 5}
\label{eq:mslrp5}
\end{equation}
which agrees with the value of the right handed slepton (c.f. Table\ref{tb:masses}).
\vskip 1.5cm
{\underline{\bf The case $\lambda_{123} \ne 0$}}
\vskip 0.2cm
As stated earlier, in this case there is always a $\tau$
particle among the decay products of the $\tilde\chi_1^0$,
and this spoils the endpoint in the OSDF lepton pair
mass distribution which would permit us to reconstruct $\tilde\chi_1^0$
and all the other sparticles further (see Fig.\ref{p5122on123} and
compare with the upper left plot of Fig.\ref{p5chi01}).
Therefore, the strategy in the case of $\rlap/R$ couplings implying the third
family of leptons must be changed, and consequently we shall {\it abandon the
direct reconstruction of $\tilde\chi_1^0$, and return to
the strategy developed in the case of conserved R parity}.
In other words, we shall discard the decay products of $\tilde\chi_1^0$
in the first stage of the reconstruction. \\
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p5122on123.eps,width=10cm }}
\caption{\small The invariant mass distribution of OSDF leptons in the case of
$\rlap/R$ coupling $\lambda_{123}=10^{-3}$ for 3 years of LHC run at low luminosity
at point 5. {\it Open histogram} - all events;
{\it shadded histogram} - background and {\it cross hatched histogram} -signal.
The arrow points to the nominal value of $\tilde \chi^0_1$ mass.
The presence of a $\tau$ lepton in the $\tilde \chi^0_1$ decay will spoil the
endpoint at the mass of $\tilde \chi^0_1$. }
\label{p5122on123}
\end{Fighere}
\vskip 1.0cm
{\it \bf Reconstruction of
$\tilde \chi^0_2 \rightarrow \tilde l^{\pm}_R + l^{\mp} \rightarrow
\tilde \chi^0_1 + l^{\mp} + l^{\pm}$
($\tilde \chi^0_1 \rightarrow \nu_{e(\mu)}+\mu^{\pm}(e^{\pm})+\tau^{\pm}$)
chain}
\vskip 0.2cm
The decay of the $\tilde \chi^0_2$ produces in this case at least
three isolated leptons.
With these leptons one can form pairs of opposite sign (OS) and different
flavors (DF) - typical for $\tilde \chi^0_1$ decay, or of the same flavor (SF)
- typical for $\tilde \chi^0_2$.
As one mentioned before, $\tilde \chi^0_2$ will produce firstly an OSSF pair
of leptons through a double 2-body decay and in a second stage,
$\tilde \chi^0_1$ will produce again one or two leptons.
In the conserved $R$ parity case one has used the OSSF combinations.
Typically, the lepton associated with $\tilde l_R$ has
a higher $p_t$ than the lepton associated with $\tilde \chi^0_1$
as it can be seen from Fig.\ref{p5123ptlep}.
In these conditions, it is useful to impose, beyond the usual cuts
(i.e. lepton multiplicity, $E_t^{miss}$, etc), a minimal difference between
the $p_t$ of the two leptons taken in the OSSF pairs.
To decrese the number of bad combinations with leptons coming from the subsequent $\tau$ decay,
one increases slightly the $p_t^{min}$ cut on leptons.
The OSSF leptons, coming from different decays, are less correlated
therefore we relax the cut in angle between them. To eliminate the
events with $\tilde \chi^0_2 \rightarrow \tilde \chi^0_1 + h^0$ we demand furthermore
that no $h^0$ has to be reconstructed in the event.
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p5123ptlep.eps,width=12cm }}
\vskip -6.0 cm
\caption{\small The $p_t$ distributions of the leptons
associated with $\tilde l_R$ or $\tilde \chi^0_1$ in the
$\tilde \chi^0_2$ decay for 3 years of LHC run at low luminosity at point 5.}
\label{p5123ptlep}
\end{Fighere}
\vskip 0.5cm
In Fig.\ref{p5123ossf}, are presented the invariant mass distributions of
OSSF lepton pairs, where the events were selected using the following critera: \\
\indent \indent $(i)$ $N_l \ge 4$, \\
\indent \indent $(ii)$ $p_t^l \geq $ 15 GeV , cos($\alpha_{l^{\pm}l^{\mp}}$) $\geq$ 0,\\
\indent \indent $(iii)$ $E_t^{miss} \ge $ 50 GeV, \\
\indent \indent $(iv)$ $\Delta p_t^l = | p_t^{l_1} - p_t^{l_2} | \geq$ 60 GeV,\\
\indent \indent $(v)$ No $b b$ pair with invariant mass in $(m_{h^0}-15,m_{h^0}+15)$ GeV.\\
The huge background comes from the bad combinations with the the leptons
produced in the $\tilde \chi^0_1$ decays.
Since the $\tilde \chi^0_2$ decays into the OSSF lepton pair via two 2-body decays,
the observed endpoint in the OSSF distribution is determined by the relation of
Eq.(~\ref{eq:mend22b})
between $m_{\tilde \chi^0_2}$, $m_{\tilde l_R}$ and $m_{\tilde \chi^0_1}$.
After the fit of the Maxwellian background and subtraction, the value of the endpoint
is estimated as :
\begin{equation}
m_{OSSF}^{meas} = 111.9 \pm 2.5 \mbox{\ GeV}
\label{eq:mOSSFp5}
\end{equation}
where the error of about 2.5 GeV is dominated by the histogram bining (i.e. statistics).
One can isolate lepton candidates coming from the double 2-body decay of
$\tilde \chi^0_2$ by a selection around this endpoint.
With the remaining
leptons one can attempt to form OS lepton pairs and further combining them
with the selected OSSF pairs to reconstruct the $\tilde \chi^0_2$.
Because of the neutrinos (always present) the $\tilde \chi^0_2$
reconstruction is not complete and therefore an endpoint will appear in this
OSSF+OS lepton invariant mass distribution depending on the $m_{\tilde \chi^0_2}$.
\vskip 0.5 cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p5123chi.eps,height=10.5cm,width=15cm }}
\caption{\small The invariant mass distribution of the OSSF lepton
pairs for 3 years of LHC run at low luminosity at point 5.
}
\label{p5123ossf}
\end{Fighere}
\vskip 0.5cm
If one applies the additional cuts to $(i \div v)$ :\\
\indent \indent $(vi)$ $m_{OSSF} \in (m^{endp}_{OSSF}-25,m^{endp}_{OSSF})$ GeV and rescaled
as explained in the introductory part of subsection 5,\\
\indent \indent $(vii)$ OS pairs constructed with leptons having $p_t^l \geq 10$ GeV,
$\Delta p_t^l \geq 20$ GeV and cos($\alpha_{OS}$) $\geq$ 0,\\
\indent \indent $(viii)$ cos($\alpha_{OSSF, OS}$) $\geq$ 0.85 ,\\
one obtains the distribution represented in Fig.\ref{p5123dchi0}.\\
One can observe an endpoint at:
\begin{equation}
m_{\tilde \chi^0_2} = 228.2 \pm 5 \mbox{\ GeV }
\label{eq:p5123chi02}
\end{equation}
by linear extrapolation of the distribution as is the case for a 5-body final state
(see ~\cite{GMSB}). This endpoint corresponds to the kinematical limit of the
$\tilde \chi^0_2$ decay, namely to $m_{\tilde \chi^0_2}$.
The error of the extrapolation is reflected in the large value of the error in
$m_{\tilde \chi^0_2}$.
\vskip 1. cm
\indent {\it \bf Reconstruction of $\tilde q_L \rightarrow \tilde \chi^0_2 + q$}
\vskip 0.2 cm
We observe that at point 5 $m_{\tilde\chi_2^0}-m_{\tilde\chi_1^0}\approx m^{endp}$
(Eq(\ref{eq:mend22b})) which means that the three-momentum of the $\tilde\chi_1^0$
is nearly zero in the $\tilde\chi_2^0$ restframe. Therefore
selecting OSSF lepton pairs near the endpoint of the distribution of Fig.\ref{p5123ossf}
one can determine $P^{\mu}_{\tilde \chi^0_2}$ using Eq.(\ref{eq:pAa})
with the value of $m_{\tilde \chi^0_2}$ from Eq(~\ref{eq:p5123chi02}).
These 4-momenta are then combined with a hard light jet not coming from a reconstructed
$W$ or $Z^0$. To the criteria $(i) \div (v)$ one adds the following :\\
\indent \indent $(vi-b)$ $m_{OSSF} \in (m^{endp}_{OSSF}-30,m^{endp}_{OSSF})$ GeV and rescaled,\\
\indent \indent $(ix)$ $p_t^{jet} \geq 400$ GeV and no invariant mass with any other light jet
in $\pm$ 15 GeV \\
\indent \indent \ \ \ \ \ \ \ around $W$,\\
\indent \indent $(x)$ cos($\alpha_{j OSSF}$) $\geq$ 0. . \\
The resulting distribution is depicted in Fig.\ref{p5123sql}.
By a gaussian fit on the peak one obtains the value:
\begin{equation}
m_{\tilde q_L}^{meas} = 684 \pm 15 \mbox{\ GeV}
\label{eq:p5123sql1}
\end{equation}
\vskip -1.0 cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p5123dchi0.eps,height=11cm,width=13cm }}
\vskip -5. cm
\caption{\small The invariant mass distribution of the OSSF+OS lepton
pairs for 3 years of LHC run at low luminosity at point 5.
Due to the presence of neutrinos (from $\tilde \chi^0_1$ decay) an endpoint
depending on $m_{\tilde \chi^0_2}$ and $m_{\tilde \chi^0_1}$ appears (see text).}
\label{p5123dchi0}
\end{Fighere}
\vskip -0.5 cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p5123sql.eps,width=9cm }}
\vskip -0.5 cm
\caption{\small The invariant mass distribution of the (jet OSSF)
pairs for 3 years of LHC run at low luminosity at point 5.
The arrow points to the nominal value of $\tilde q_L$ mass. The mean value is obtained by
a local gaussian fit on the main peak.}
\label{p5123sql}
\end{Fighere}
\vskip 0.5cm
$m_{\tilde q_L}$ can also be determined by combining the (OSSF+OS) pairs selected near
the $m_{\tilde \chi^0_2}$ endpoint of Fig.\ref{p5123dchi0}, whose 4-momenta are approximately equal to
$P^{\mu}_{\tilde \chi^0_2}$ according to Eq.(~\ref{eq:pAa}), with a hard light jet
not coming from a reconstructed $W$ or $Z^0$. For that one will use
the cuts $(i \div ix)$ supplemented by:\\
\indent \indent $(xi)$ $m_{\tilde \chi^0_2} \in (m^{endp}_{OSSF,OS}-50,m^{endp}_{OSSF,OS})$ GeV,\\
\indent \indent $(xii)$ cos($\alpha_{j OSSF-OS}$) $\geq$ 0 . \\
The invariant mass distribution is shown in Fig.~\ref{p5123sql2}. One obtains the value:
\begin{equation}
m_{\tilde q_L}^{meas} = 686 \pm 12 \mbox{\ GeV.}
\label{eq:p5123sql2}
\end{equation}
The agreement between the results of (~\ref{eq:p5123sql1}) and (~\ref{eq:p5123sql2}) justifies the
procedure that we have applied in obtaining the mass of $\tilde \chi^0_2$.
\vskip 0.5 cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p5123sql2.eps,width=10cm }}
\caption{\small
The invariant mass distribution of the (jet OSSF-OS) pairs
for 3 years of LHC run at low luminosity at point 5.
The arrow points to the nominal value of $\tilde q_L$ mass.
}
\label{p5123sql2}
\end{Fighere}
\vskip 0.5cm
\vskip 1.0 cm
\indent {\it \bf Reconstruction of $h^0 \rightarrow b + \bar b$}
\vskip 0.2 cm
The reconstruction of the $h^0$ can be carried out almost with the same precision
as it was the case of the coupling $\lambda_{122}$.
However, the applied isolation criteria reject more events due to the finite size
of the $\tau$ jets as compared to the case of $\lambda_{122}$.
This can be seen in the $b \bar b$ invariant mass
distributions which should be independent of the $\rlap/R$ coupling type.
Following the same procedure as in the case $\lambda_{122}$, with the selection
criteria slightly modified ;\\
\indent \indent $(xiii)$ 15 GeV $\leq p_t^b \leq$ 300 GeV,\\
\indent \indent $(xiv)$ cos($\alpha_{bb}$) $\geq$ 0.4 GeV,\\
one obtains the distribution shown in Fig.\ref{p5123h0}.
After the subtraction of the Maxwellian background and the
gaussian fit of the peak one gets:
\begin{equation}
m_{h^0}^{meas} = 94.3 \pm 1.5 \mbox{\ GeV}
\label{eq:p5123h0mass}
\end{equation}
This number includes the systematic error, mainly due to the incertitude on the energy
scale of $b$ jets.
\vskip -1.0 cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p5123h0.eps,width=16cm }}
\caption{\small The invariant mass distribution of the ($bb$)
pairs for 3 years of LHC run at low luminosity at point 5.
}
\label{p5123h0}
\end{Fighere}
\vskip 1.0 cm
\subsection{The LHC point No3}
\vskip 1.0 cm
The salient feature of this point w.r.t. the other two
is that the SUGRA parameters $m_0$ and $m_{1/2}$ are much lower, hence
all sparticle masses are lighter and the production
cross sections are highly increased. As a consequence we could
generate only a fraction of the events one would be able to detect
with the ATLAS detector in one year even with low luminosity.
In addition, due to the $sign\mu=-1$ and
$m_{\tilde\chi^0_2}-m_{\tilde\chi^0_1}\simeq 52$ GeV,
the second lightest neutralino decays into OS and SF lepton pairs and not
into a Higgs ($h^0$, with a mass $\sim$ 70 GeV).
Therefore in this point one cannot determine the Higgs mass.
The number of leptons in the final state being higher, the
combinatorial background is considerably larger in the
channels where the particle reconstruction involves leptons. Furthermore,
as we precised before, $\tilde \chi^0_1$ can decay
(through $\rlap/R$ $\lambda$-type couplings) in lepton pairs of different flavors as well
as of same flavors. The consequence is a much more complicated structure
in the invariant mass distribution of the OSSF lepton pairs than in the case of conserved
R parity.
Another feature of point 3 is that in the decay chain (*) the dominant
decay products of the gluino are the $\tilde b_1$ and the $b$. Since in the
decay of the $\tilde b_1$ the gaugino ($\tilde\chi_2^0$ or $\tilde\chi_1^{\pm}$)
is accompanied again by a $b$-quark, there is a large number of
$b$-quarks produced in each event.
In the case of conserved $R$ parity~\cite{SUGRA3} the observables which can be used
for the determination of the SUGRA parameters are {\it the functions}
$m_{\tilde g}(m_{\tilde \chi^0_2})$, $m_{\tilde b_1}(m_{\tilde \chi^0_2})$
and the {\it difference} $\Delta m_{\tilde\chi^0} = m_{\tilde\chi_1^0}-m_{\tilde\chi_2^0}$
(from the invariant mass distribution of the OS and SF lepton pairs:
$\tilde\chi_2^0\longrightarrow (l^+l^-) +\tilde\chi_1^0 $
- see the decay chain (*) and Equ.~(\ref{eq:mAa})).
In the $\rlap/R$ case, one can develop two strategies depending on the $\lambda$ type coupling :\\
\indent (1) \ \ For $\lambda$ couplings not producing $\tau$ jets in the LSP decays
(i.e. $\lambda_{122}$) one will {\it directly} reconstruct the $\tilde \chi^0_1$ as in the
other LHC points. The consequence of this new information is a better fit of the SUGRA parameters
comparing with the $R$ conserved case;\\
\indent (2) \ \ For $\lambda$ couplings producing $\tau$ jets in the LSP decays
(i.e. $\lambda_{123}$) one returns to the strategy developed in the $R$ conserved case.
As in the LHC points 1 and 5, we will present distinctively the two
representative cases: $\lambda_{122}$ and $\lambda_{123}$.
\vskip 0.8cm
{\bf \large \it \underline{The case $\lambda_{122} \ne 0$}}
\vskip 0.2cm
The analysis is focused firstly on the leptons. With these, one can form OS pairs
of the same flavor (SF) or different flavors (DF).
Due to the facts that:\\
\indent - $\tilde \chi^0_1$ has a significantly higher production rate than the
$\tilde \chi^0_2$ and \\
\indent - only $\tilde \chi^0_1$ produces $OSDF$ lepton pairs,\\
the OSDF pairs will tag much better their origine than the OSSF ones.
Therefore, in the first stage one will reconstruct the invariant mass distribution of OSDF
leptons. This gives the $\tilde \chi^0_1$ mass through the same type of
endpoint structure as in the other LHC points. The main background in this distribution
is combinatorial. One will select OSDF lepton pairs in the
neighbourhood of this endpoint.
With the remaining leptons (originating with a higher probability from
$\tilde \chi^0_2$, but also fom $\tilde \chi^0_1$) one will form OS lepton pairs.
The endpoint in their invariant mass distribution determines
$\Delta m_{\tilde \chi^0} = m_{\tilde \chi^0_2}-m_{\tilde \chi^0_1}$ .
Selecting OS pairs near this latter endpoint and combining them with the
$\tilde \chi^0_1$ candidates (from the OSDF distribution) one can completely reconstruct
$\tilde \chi^0_2$. Afterwards, combining $\tilde \chi^0_2$ candidates with one and
two $b$ jets one will reconstruct $\tilde b_1$ and $\tilde g$.\\
For this analysis one uses the following global cuts :\\
\indent \indent ($i$)\ \ $N_l \geq 4$, and $p_t^{e, \mu} \geq$ 10 GeV,\\
\indent \indent ($ii$) $E_t^{miss} \geq 50$ GeV.\\
\vskip 0.6 cm
{\it \bf Reconstruction of
$\tilde \chi^0_1 \rightarrow \nu_{e(\mu)}+\mu^{\pm}(e^{\pm})+\mu^{\pm}$}
\vskip 0.2cm
Using the global cuts ($i \div ii$) and :\\
\indent \indent ($iii$) cos($\alpha_{OSDF}$) $\geq 0.85$,\\
one obtains the invariant mass distribution of OSDF leptons as in Fig.\ref{p3122osdf},
where the background is mainly combinatorial.
After the fit of the Maxwellian background, subtraction and the polinomial fit of the
resulting endpoint one obtains the $\tilde \chi^0_1$ mass :
\begin{equation}
m_{\tilde \chi^0_1}^{meas} = 44.8^{+0.1}_{-0.2} \mbox{\ GeV}
\label{eq:p3122chi01mass}
\end{equation}
\vskip -0.5 cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p3122osdf.eps,width=14cm }}
\caption{\small The invariant mass distribution of the OSDF lepton
pairs for 1 year of LHC run at low luminosity at point 3 in the case of
$\rlap/R$ coupling $\lambda_{122}=10^{-3}$.
}
\label{p3122osdf}
\end{Fighere}
\vskip 0.5 cm
\vskip 0.6 cm
{\it \bf Reconstruction of
$\tilde \chi^0_2 \rightarrow \tilde \chi^0_1 + \mu^{\pm}(e^{\pm})+\mu^{\mp}(e^{\mp})$}
\vskip 0.2cm
The $\tilde \chi^0_1$ candidates are identified by requiring OSDF lepton pairs with :\\
\indent \indent ($iv$)
$m(OSDF) \in (m_{\tilde \chi^0_1}^{meas}-10, m_{\tilde \chi^0_1}^{meas})$ GeV,\\
with rescaled 4-momenta according to Eq.(\ref{eq:pAa}).
One then selects leptons produced directly from the 3-body decay
of $\tilde \chi^0_2$ among the remaining OS lepton pairs demanding that the angle
between the leptons ($\alpha_{OS}$) satisfies :\\
\indent \indent ($v$) cos($\alpha_{OS}$) $\geq 0.5$ .\\
The invariant mass distribution of these OS pairs (Fig.\ref{p3122os}) has a complex structure
due to the fact that,
beside the combinatorial background (Maxwellian type with a long tail) there is also a contribution from
the $\tilde \chi^0_1$ decay itself (in fact the contribution is 50\% of its branching ratio and it has
the shape comparable with that of OSDF distribution). Therefore the OS distribution shows
a sharp edge at about 52 GeV, correspondig to the mass difference
$m_{\tilde \chi^0_2} - m_{\tilde \chi^0_1}$
and a second one, less pronounced around 45 GeV, corresponding to the $\tilde \chi^0_1$
decay products. The tail beyond 52 GeV is due to the combinatorial background
and we have removed it. The polinomial fit of the higher endpoint gives the value:
\begin{equation}
m^{meas}_{OS} = m_{\tilde \chi^0_2}-m_{\tilde \chi^0_1} = 53.3 \pm 0.6 \mbox{\ GeV}
\label{eq:p3122osmass}
\end{equation}
\vskip -0.5 cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p3122os.eps,width=14cm }}
\caption{\small The invariant mass distribution of the remaining OS lepton
pairs, after the removal of $\tilde \chi^0_1$ candidates. The distributions corresponds to
1 year of LHC run at low luminosity at point 3 in the case of
$\rlap/R$ coupling $\lambda_{122}=10^{-3}$.
For explanations see text.}
\label{p3122os}
\end{Fighere}
\vskip 0.5 cm
As a next step we apply the cuts ($i \div v$) and :\\
\indent \indent ($vi$)
$m(OS) \in (m_{OS}^{meas}-10, m_{OS}^{meas})$ GeV,\\
in order to reconstruct the $\tilde \chi^0_2$ by combining the OSDF and the remainig OS lepton pairs.
We require that the angle between the two pairs should be small :\\
\indent \indent ($vii$) cos($\alpha_{OSDF, OS}$) $\geq 0.8$. \\
The obtained mass distribution is shown in the Fig.\ref{p3122chi02}.
The gaussian fit around the peak results in the value :
\begin{equation}
m^{meas}_{\tilde \chi^0_2} = 96.7 \pm 0.2 \mbox{\ GeV}
\label{eq:p3122chi02mass}
\end{equation}
Out of the three quantities defined by the Eq.(\ref{eq:p3122chi01mass}), (\ref{eq:p3122osmass}) and
(\ref{eq:p3122chi02mass}) we use two independent ones
(Eq.(\ref{eq:p3122chi01mass}) and (\ref{eq:p3122chi02mass})) ,
for the determination of the SUGRA parameters.
\vskip -0.5 cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p3122chi02.eps,height=9.0cm,width=12.0cm }}
\caption{\small The invariant mass distribution OSDF $+$ OS lepton pairs.
The distributions correspond to 1 year of LHC run at low luminosity at point 3.
For explanations see text.}
\label{p3122chi02}
\end{Fighere}
\vskip 0.8cm
{\it \bf Reconstruction of the chain
$\tilde g \rightarrow \tilde b_1 + b \rightarrow \tilde \chi^0_2 + b + b$}
\vskip 0.2cm
The configuration of the sparticle mass spectrum (see Table\ \ref{tb:masses}) and the decay branching
ratios for that point, permit us to make the following remarks on the production of $b$ jets:
\indent (1) practically all $b$ jets originate from the decay chain of $\tilde g$ (with a small
fraction coming also from $\tilde t_1$ and/or $t$),\\
\indent (2) the $b$ jets from the first decay of $\tilde g$ are very soft comparing to
those produced in the second decay (and associated with $\tilde \chi^0_2$),
as it can be seen in the Fig.\ref{p3123ptbj}. Correspondingly we label the $b$ jets as follows :\\
\indent \indent ($viii$) {\it hard} $b$ jets if : $ p_t^b \geq$ 50 GeV ;\\
\indent \indent ($ix$) \ \ {\it soft} $b$ jets if : 10 GeV $\leq p_t^b \leq$ 50 GeV .\\
Selecting $\tilde \chi^0_2$ candidates around the $\tilde \chi^0_2$ peak from Fig.\ref{p3122chi02} :\\
\indent \indent ($x$) \ \ $m(OSDF,OS) \in (m^{peak}-15,m^{peak}+15)$ GeV,\\
one performs a first reconstruction of the $\tilde g$ mass as follows. First of all, one selects pairs
of $b$ jets with all $b$ jets passing the cut $p_t \geq$ 10 GeV and in each pair we identify the
{\it hard} and the {\it soft} jet. Next we combine the $\tilde \chi^0_2$, $b_{hard}$ and
$b_{soft}$ 4-momenta if :\\
\indent \indent ($xi$) \ cos($\alpha_{\tilde \chi^0_2 b_{hard}}$) $\geq 0$ , and\\
\indent \indent ($xii$) cos($\alpha_{(\tilde \chi^0_2 b_{hard}),b_{soft}}$) $\geq 0.5$. \\
These cuts are justified by the distributions shown in Fig.\ref{p3122rawgl}a and \ref{p3122rawgl}b.
\vskip -1.5 cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p3123ptbj.eps,height=10.0cm,width=14cm }}
\caption{\small The $p_t$ distribution of $b$ jets in point 3 for 1 year of LHC run at low
luminosity after applying the global cuts $N_l \geq 4$, $E_t^{miss} \geq 50$ GeV and
$p_t^b \geq 10$ GeV. One can clearely distinguish the $b$ jets from $\tilde g$
and from $\tilde b_1$.}
\label{p3123ptbj}
\end{Fighere}
\begin{Fighere}
\hspace{-1.0cm}
\centering
\mbox{\epsfig{file=./p3122rglucos.eps,height=8.0cm }\ \
\epsfig{file=./p3122rglumass.eps,height=8.0cm }}
\vskip -0.5cm
\caption{\small The first reconstruction of the gluino mass in point 3, for
$\rlap/R$ coupling $\lambda_{122}=10^{-3}$ and after 1 year of LHC run at low
luminosity (plot c). Also shown are the angle distributions between the
$\tilde \chi^0_2$ and the hard $b$-jets (plot a) and between the ($\tilde \chi^0_2$, $b_{hard}$)
pairs and soft $b$-jets (plot b). For further comments see text.}
\label{p3122rawgl}
\end{Fighere}
\vskip 0.5 cm
The resulting mass distribution is represented in Fig.\ref{p3122rawgl}c.
For the $\tilde b_1$ reconstruction, one selects the $\tilde \chi^0_2$ candidates
according to ($x$) and the {\it hard} $b$ jets (see ($viii$)) around the
$\tilde g$ peak of the Fig.\ref{p3122rawgl}c :\\
\indent \indent ($xiii$) $m_{\tilde g} \in (m^{peak}-10, m^{peak}+10)$ GeV, requiring this time \\
\indent \indent ($xiv$) \ cos($\alpha_{\tilde \chi^0_2 b_{hard}}$) $\geq 0.5$.\\
The result is represented in Fig.\ref{p3122sb1}. After the gaussian fit around the peak one
obtains the value :
\begin{equation}
m_{\tilde b_1}^{meas} = 276.6 \pm 3.0 \mbox{\ GeV}
\label{eq:p3122sb1mass}
\end{equation}
\vskip -0.5 cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p3122sb1.eps,width=11cm }}
\caption{\small The invariant mass distribution of $\tilde \chi^0_2$ and $b_{hard}$ jets.
The distribution corresponds to 1 year of LHC run at low
luminosity. For the selection criteria see text.}
\label{p3122sb1}
\end{Fighere}
\vskip 0.5 cm
This permits us to refine the reconstruction of the $\tilde g$ by tagging it with the $\tilde b_1$.
We combine the $\tilde \chi^0_2$ candidates (cut ($xi$)) with a {\it hard} $b$ (cut ($viii$))
requiring an invariant mass within the $\tilde b_1$ mass window: \\
\indent \indent ($xv$)
$m_{\tilde \chi^0_2 b_{hard}} \in (m_{\tilde b_1}^{meas}-10, m_{\tilde b_1}^{meas}+10)$ GeV.\\
We add subsequently a {\it soft} $b$ jet with the same angular correlation as ($xii$) to obtain
the final gluino reconstruction shown in the Fig.\ref{p3122gl}.
The gaussian fit around the peak gives the measured value :\\
\begin{equation}
m_{\tilde g}^{meas} = 301.1 \pm 3.0 \mbox{\ GeV .}
\label{eq:p3122glmass}
\end{equation}
\vskip -0.5 cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p3122fglu.eps,width=9cm }}
\caption{\small The final reconstruction of the $\tilde g$ after
having selected the ($\tilde \chi^0_2$, $b_{hard}$) pairs around the reconstructed
$\tilde b_1$ mass peak. The distribution corresponds to 1 year of LHC run at low
luminosity.}
\label{p3122gl}
\end{Fighere}
\vskip 0.8cm
{\bf \large \it \underline{The case $\lambda_{123} \ne 0$}}
\vskip 0.5cm
{\it \bf Reconstruction of
$\tilde \chi^0_2 \rightarrow \tilde \chi^0_1 + \mu^{\pm}(e^{\pm})+\mu^{\mp}(e^{\mp})$}
\vskip 0.2cm
Comparing to the case $\lambda_{122}$ the presence of $\tau$ jets in the decay products of
$\tilde \chi^0_1$ spoils the endpoint structure in the OSDF lepton pair distribution and
therefore the direct reconstruction of the $\tilde \chi^0_1$.
Although the electrons and muons from the $\tau$ jets increase the combinatorial background
of the OSSF lepton pairs, due the fact that they are softer than those from the 3-body decay of
the $\tilde \chi^0_2$ they do not spoil the endpoint structure.
We select events with the usual cuts ($i \div ii$) and OSSF pairs by requiring :\\
\indent \indent ($xvi$) cos($\alpha_{OSSF}$) $\geq$ 0.5.\\
After a fit of the Maxwellian tail, subtraction and the fit of the result with a polinomial
(see Fig.\ref{p3123dchi0}), one obtains using Eq.(\ref{eq:mAa}) :
\begin{equation}
m_{OSSF}^{meas}=m_{\tilde \chi^0_2}-m_{\tilde \chi^0_1} = 52.9^{+0.1}_{-0.3} \mbox{\ GeV .}
\label{eq:p3123ossfmass}
\end{equation}
\vskip 0.8cm
{\it \bf Reconstruction of the chain
$\tilde g \rightarrow \tilde b_1 + b \rightarrow \tilde \chi^0_2 + b + b$}
\vskip 0.2cm
Since the $\lambda$ couplings do not affect the $b$ jets (if we neglect the difference in the
reconstruction and tagging efficiency due to the larger size of a $\tau$ jet comparing to a lepton)
the classification of $b$ jets in {\it soft} and {\it hard} types is done by the same
criteria as in the case $\lambda_{122}$.
The reconstruction of this chain follows the same treatement
as in the case $\lambda_{122}$ with the notable difference that this time one cannot
directly reconstruct the $\tilde \chi^0_2$. Instead, we have to assume a mass value for the
$\tilde \chi^0_2$ in order to obtain its 4-momentum at the endpoint of the OSSF lepton pair
mass distribution according to the Eq.(\ref{eq:pAa}). We have taken $m_{\tilde \chi^0_2}=97$ GeV.
\vskip -1.0 cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p3123dchi0.eps,height=13.0cm,width=15.0cm }}
\caption{\small The invariant mass distribution of the OSSF lepton pairs
for 1 year of LHC run at low luminosity at point 3 in the case of
$\rlap/R$ coupling $\lambda_{123}=10^{-3}$ .}
\label{p3123dchi0}
\end{Fighere}
\vskip 0.5 cm
The cuts used in this case are the following :\\
\indent \indent ($x$)$^*$ \ \ $m(OSSF) \in (m^{endp}_{OSSF}-10, m^{endp}_{OSSF})$ GeV,
for the $\tilde \chi^0_2$ candidates,\\
and for the first reconstruction of the gluino :\\
\indent \indent ($xi$)$^*$ \ cos($\alpha_{\tilde \chi^0_2 b_{hard}}$) $\geq$ 0.5, \\
\indent \indent ($xii$)$^*$ cos($\alpha_{\tilde \chi^0_2 b_{soft}}$) $\geq$ 0.5. \\
The invariant mass distribution of $\tilde \chi^0_2 b_{hard} b_{soft}$ pairs
is shown in Fig.\ref{p3123rawgl}.
Selecting the {\it hard} $b$ jets if around the gluino peak :\\
\indent \indent ($xiii$)$^*$ $m_{\tilde g} \in (m^{peak}-10, m^{peak}+10)$ GeV and,\\
\indent \indent ($xiv$)$^*$ \ cos($\alpha_{\tilde \chi^0_2 b_{hard}}$) $\geq 0.5$.\\
one obtains the $\tilde b_1$ mass peak as shown in the Fig.\ref{p3123sb1}.
The gaussian fit around the peak gives :
\begin{equation}
m_{\tilde b_1}^{meas}(97) = 277.5 \pm 3.0 \mbox{\ GeV}
\label{eq:p3123sb1mass}
\end{equation}
The 97 inside the brackets indicates that this value was obtained assuming
$m_{\tilde \chi^0_2} = $ 97 GeV.
\vskip -0.5 cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p3123rawgl.eps,height=8.0cm,width=12cm }}
\caption{\small The first reconstruction of the gluino in point 3, for
$\rlap/R$ coupling $\lambda_{123}=10^{-3}$ and after 1 year of LHC run at low
luminosity .}
\label{p3123rawgl}
\end{Fighere}
\vskip 0.5 cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p3123sb1.eps,width=11cm }}
\caption{\small The reconstruction of the $\tilde b_1$ selecting the $b$ jet around the
$\tilde g$ peak. The distribution corresponds to 1 year of LHC run at low
luminosity.}
\label{p3123sb1}
\end{Fighere}
\vskip 0.5 cm
Finally, selecting the {\it hard} $b$ jets and the OSSF pairs around the $\tilde b_1$ peak :\\
\indent \indent ($xv$)$^*$
$m_{\tilde \chi^0_2 b_{hard}} \in (m_{\tilde b_1}^{meas}-10, m_{\tilde b_1}^{meas}+10)$ GeV,\\
we combine them with the {\it soft } $b$ jets with the same angle cuts as above
(($xi$)$^*$ and $xii$)$^*$). The resulting mass distribution is shown in the Fig.\ref{p3123gl}.
The gaussian fit around the peak gives the measured value :\\
\begin{equation}
m_{\tilde g}^{meas}(97) = 301.1 \pm 3.5 \mbox{\ GeV .}
\label{eq:p3123glmass}
\end{equation}
\vskip -0.5 cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p3123gl.eps,width=12.0cm }}
\caption{\small The final reconstruction of the $\tilde g$ after having selected the OSSF lepton
pairs and {\it hard} $b$ jets around the $\tilde b_1$ peak.
The distribution corresponds to 1 year of LHC run at low luminosity. }
\label{p3123gl}
\end{Fighere}
\vskip 0.5cm
We have repeated the above analysis for several
different values of $m_{\tilde \chi^0_2}$. The result is
shown in Fig.~\ref{fg:p3chi02theor}. Using a linear fit one obtains the following
expressions :
\begin{equation}
m_{\tilde b_1}^{meas}(m_{\tilde \chi^0_2}) = m_{\tilde b_1}^{meas}(97) +
\theta_{\tilde b_1}(m_{\tilde \chi^0_2}-97) \mbox{ GeV}
\label{eq:p3123sb1_var_mass}
\end{equation}
\begin{equation}
m_{\tilde g}^{meas}(m_{\tilde \chi^0_2}) = m_{\tilde g}^{meas}(97) +
\theta_{\tilde g}(m_{\tilde \chi^0_2}-97) \mbox{ GeV}
\label{eq:p3123gl_var_mass}
\end{equation}
with $\theta_{\tilde b_1} = 1.48 \pm 0.52$ and $\theta_{\tilde g} = 1.55 \pm 0.55$, repectively.
\vskip -0.5cm
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p3chi02exp.eps,width=11.0cm }}
\caption{\small The dependence of $m_{\tilde b_1}$ (full squares) and $m_{\tilde g}$ (full circles)
as functions of $m_{\tilde \chi^0_2}$. The arrow points to the nominal value of
$m_{\tilde \chi^0_2}=97$ GeV. The factors $\theta_{\tilde b_1}$ and $\theta_{\tilde g}$ (see text)
are extracted by a linear fit.}
\label{fg:p3chi02theor}
\end{Fighere}
\vskip 1.0 cm
\subsection{Determination of the SUGRA parameters}
\vskip 0.5 cm
In subsection 2.2 we have already reviewed the dependence of the sparticle masses
as functions of the model parameters $m_0$, $m_{1/2}$, $A_0$, tan$\beta$ and sign$\mu$.
Practically all sparticle masses are sensitive to $m_{1/2}$. $m_0$ drives mainly the sfermion masses.
sign$\mu$ affects the gaugino branching ratios
(i.e. $\chi^0_2 \rightarrow \chi^0_1 + l^{+} + l^{-}$ versus $\chi^0_2 \rightarrow \chi^0_1 + h^0$),
but also the sparticle mass spectrum. On the other hand, the observables are practically
independent of $A_0$ (c.f. Fig~\ref{fg:p5alla0fit} and Table~\ref{tb:slope_p5a0}).
An obvious way to determine the above parameters consists of finding the minimum of :
\begin{equation}
\begin{tabular}{l}
$\chi^2(m_0,m_{1/2},tan\beta,A_0) =$ \\
\\
$\sum_{\alpha,\beta=1}^{N_{obs}} [O^{th}_{\alpha}(m_0,m_{1/2},tan\beta,A_0) - O^{meas}_{\alpha}]
\sigma^{-1}_{\alpha \beta}[O^{th}_{\beta}(m_0,m_{1/2},tan\beta,A_0) - O^{meas}_{\beta}]$
\end{tabular}
\label{eq:chisquare}
\end{equation}
where $O^{th}_{\alpha}(m_0,m_{1/2},tan\beta,A_0)$ represents the theoretical evolution of the
observable $\alpha$ in function of the SUGRA parameters, $O^{meas}_{\alpha}$ is the measured value
of the same observable and $\sigma_{\alpha \beta}$ is the covariance matrix
\footnote{In general $O^{meas}_{\alpha}$ are not independent, however at the present time we neglect the
offdiagonal elements of $\sigma_{\alpha \beta}$.} .\\
The minimum of the $\chi^2$ can be found by scanning through the entire parameter
space \cite{Froid_LHCC}.
The overall minimum is practically insensitive to $A_0$. The sign$\mu$ parameter can be
determined unambiguosly \cite{Froid_LHCC}. Once the overall minimum is found one can determine the
error on the parameters at the minimum by taking the square root of the diagonal elements of the
matrix $\Delta_{ij}$ :
\begin{equation}
\Delta_{ij} = \Bigl[ \sum_{\alpha ,\beta=1}^{N_{obs}}
A^{T}_{i\alpha}(\sigma_{\alpha \beta })^{-1} A_{\beta j} \Bigr]^{-1}, \ \ \ \
A_{\alpha i} = \frac{\partial O^{th}_{\alpha}}{\partial (p_i)},
\ \ p_i \equiv m_0, m_{1/2}, tan \beta, A_0, \theta_{\tilde g}, \theta_{\tilde b_1}
\label{eq:paramerror}
\end{equation}
Table~\ref{tb:values} summarizes all the measured values and errors of the chosen observables
for all studied cases. For point 3 we have added the $h^0$ mass determined by an
independent measurement ($h^0 \rightarrow \gamma \ \gamma$) with an estimated error of 1 GeV.
As one can see from this Table in most of the cases the models are overconstrained : one disposes
of more measured quantities than the number of parameters to be determined. Comparing the measured
values of the observables with the theoretical ones (see Table~\ref{tb:masses}) one can see that
the errors are slightly overestimated.
$O^{th}(p_i)$ are the theoretical dependencies of the quantitities $O^{meas}$ collected in
Table~\ref{tb:values}. However, there is an exception at point 3 (for $\lambda_{123} \neq 0$)
where, according to Eqs.(\ref{eq:p3123sb1_var_mass}) and (\ref{eq:p3123gl_var_mass}) one has :
\begin{equation}
\begin{tabular}{l}
$O^{th}_{\tilde g}(p_i)=m_{\tilde g}(p_i)-
\theta_{\tilde g}(m_{\tilde \chi^0_2}(p_i)-97)$, \hskip 0.5cm
$p_i=m_0, m_{1/2}, A_0, tan\beta$ \\
$O^{th}_{\tilde b_1}(p_i)=m_{\tilde b_1}(p_i)-
\theta_{\tilde b_1}(m_{\tilde \chi^0_2}(p_i)-97)$\\
\end{tabular}
\label{fitwithchi02}
\end{equation}
In this case we have to add two more terms to Eq.(\ref{eq:chisquare}), namely :
\begin{equation}
\chi^2 \longrightarrow \chi^2 +
\Bigl(\frac{\theta_{\tilde g} - \theta^{m}_{\tilde g}}{\sigma_{\theta_{\tilde g}}}\Bigr)^2 +
\Bigl(\frac{\theta_{\tilde b_1} - \theta^{m}_{\tilde b_1}}{\sigma_{\theta_{\tilde b_1}}}\Bigr)^2
\label{newchisquare}
\end{equation}
and minimize this expression by varying also $\theta_{\tilde g}$ and $\theta_{\tilde b_1}$.
As one can see the errors on $\theta_{\tilde g}$ and $\theta_{\tilde b_1}$ do not contribute
to the errors of the other parameters in the first approximation.
As an example, in the Figs.~\ref{fg:p5allm0fit} $\div$ \ref{fg:p5alltanbfit} we have represented
the dependence of the observables on $m_0$, $m_{1/2}$, $A_0$ and tan$\beta$, respectively, in point 5.
Each histogram is fitted with a polinomial in a domain $\pm 50$ GeV - for $m_0$, $m_{1/2}$ and $A_0$ and
$\pm 0.5$ - for tan$\beta$, around the nominal values of point 5. The first derivatives of these
functions are used to obtain $\Delta_{ij}$ of Eq.(\ref{eq:paramerror}).
The final results are compiled in the form of relative errors in Table~\ref{tb:sugrafitp1}, taking
$A_0$ fixed at its nominal value of each point.
The obtained precision on the SUGRA parameters are in general higher than in the case of
conserved R parity \cite{Froid_LHCC}. The reason is that here we were able to determine more
observables and usually in a more direct way through the reconstruction of the LSP.
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p5allm0.eps,width=16.0cm }}
\caption{\small The dependence of $m_{\tilde \chi^0_1}$, $m_{\tilde l_R}$, $m_{\tilde q_R}$,
$m_{\tilde q_L}$, $m_{\tilde h^0}$ and $M^{end}_{OSSF}$ on $m_0$ smoothed by a polynomial fit.
All other parameters are fixed at those of point 5.}
\label{fg:p5allm0fit}
\end{Fighere}
\vskip 1.0 cm
\begin{Tabhere}
\centering
\caption{\small The slopes, $\frac{\partial O^{th}_{\alpha}}{\partial m_0}$,
of the observables at the
nominal value of point 5 : $m_0=100$ GeV.}
\vskip 0.5cm \hskip -0.8cm
\begin{tabular}{|c|c|c|c|c|c|} \hline\hline
$\tilde \chi^0_1$ & $\tilde l^{\pm}_R$ & $\tilde q_R$ & $\tilde q_L$ & $\tilde h^0$ & $M^{end}_{OSSF}$ \\ \hline\hline
0.001834 & 0.634610 & 0.15894 & 0.13929 & 0.001954 & 0.29879 \\ \hline\hline
\end{tabular}
\label{tb:slope_p5m0}
\end{Tabhere}
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p5allm12.eps,width=16.0cm }}
\caption{\small The dependence of $m_{\tilde \chi^0_1}$, $m_{\tilde l_R}$, $m_{\tilde q_R}$,
$m_{\tilde q_L}$, $m_{\tilde h^0}$ and $M^{end}_{OSSF}$ on $m_{1/2}$ smoothed by a polynomial fit.
All other parameters are fixed at those of point 5.}
\label{fg:p5allm12fit}
\end{Fighere}
\vskip 1.0 cm
\begin{Tabhere}
\centering
\caption{\small The slopes, $\frac{\partial O^{th}_{\alpha}}{\partial m_{1/2}}$,
of the observables at the
nominal value of point 5 : $m_{1/2}=300$ GeV.}
\vskip 0.5cm \hskip -0.8cm
\begin{tabular}{|c|c|c|c|c|c|} \hline\hline
$\tilde \chi^0_1$ & $\tilde l^{\pm}_R$ & $\tilde q_R$ & $\tilde q_L$ & $\tilde h^0$ & $M^{end}_{OSSF}$ \\ \hline\hline
0.452232 & 0.28396 & 1.928 & 2.0258 & 0.037934 & 0.293516 \\ \hline\hline
\end{tabular}
\label{tb:slope_p5m12}
\end{Tabhere}
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p5alla0.eps,width=16.0cm }}
\caption{\small The dependence of $m_{\tilde \chi^0_1}$, $m_{\tilde l_R}$, $m_{\tilde q_R}$,
$m_{\tilde q_L}$, $m_{\tilde h^0}$ and $M^{end}_{OSSF}$ on $A_0$ smoothed by a polynomial fit.
All other parameters are fixed at those of point 5.}
\label{fg:p5alla0fit}
\end{Fighere}
\vskip 1.0 cm
\begin{Tabhere}
\centering
\caption{\small The slopes, $\frac{\partial O^{th}_{\alpha}}{\partial A_0}$,
of the observables at the
nominal value of point 5 : $A_0=300$ GeV.}
\vskip 0.5cm \hskip -0.8cm
\begin{tabular}{|c|c|c|c|c|c|} \hline\hline
$\tilde \chi^0_1$ & $\tilde l^{\pm}_R$ & $\tilde q_R$ & $\tilde q_L$ & $\tilde h^0$ & $M^{end}_{OSSF}$ \\ \hline\hline
-0.0008018 & 0.00002584 & 0.0018422 & 0.017902 & -0.003919 & -0.001673 \\ \hline\hline
\end{tabular}
\label{tb:slope_p5a0}
\end{Tabhere}
\begin{Fighere}
\centering
\mbox{\epsfig{file=./p5alltanb.eps,width=16.0cm }}
\caption{\small The dependence of $m_{\tilde \chi^0_1}$, $m_{\tilde l_R}$, $m_{\tilde q_R}$,
$m_{\tilde q_L}$, $m_{\tilde h^0}$ and $M^{end}_{OSSF}$ on tan$\beta$ smoothed by a polynomial fit.
All other parameters are fixed at those of point 5.}
\label{fg:p5alltanbfit}
\end{Fighere}
\vskip 1.0 cm
\begin{Tabhere}
\centering
\caption{\small The slopes, $\frac{\partial O^{th}_{\alpha}}{\partial tan\beta}$,
of the observables at the
nominal value of point 5 : tan$\beta=2.1$ .}
\vskip 0.5cm \hskip -0.8cm
\begin{tabular}{|c|c|c|c|c|c|} \hline\hline
$\tilde \chi^0_1$ & $\tilde l^{\pm}_R$ & $\tilde q_R$ & $\tilde q_L$ & $\tilde h^0$ & $M^{end}_{OSSF}$ \\ \hline\hline
-1.101242 & 1.802616 & 2.99144 & 2.37354 & 11.1297 & -3.14764 \\ \hline\hline
\end{tabular}
\label{tb:slope_ptanb}
\end{Tabhere}
\newpage
\begin{Tabhere}
\centering
\caption{\small The measured values of observables in SUGRA points 1, 3 and 5 for each $\rlap/R$
coupling considered. All the values are in GeV.}
\vskip 0.5cm \hskip -0.3cm
\begin{tabular}{|c|c|c|c|c|c|} \hline\hline
$O^{meas} \pm \delta O^{meas}$ & \multicolumn{1}{c|}{Point 1} & \multicolumn{2}{c|}{Point 3} & \multicolumn{2}{c|}{Point 5} \\ \cline{2-6}
&$\lambda_{122}$ & $\lambda_{122}$ & $\lambda_{123}$ & $\lambda_{122}$ & $\lambda_{123}$ \\ \hline\hline
$m_{\tilde g}\pm
\sigma (m_{\tilde g})$ & & 301.1$\pm$3.0 & 301.1$\pm$3.5$^{3)}$ & & \\ \hline
$m_{\tilde q_R}\pm
\sigma (m_{\tilde q_R})$ & 932$\pm$20 & & & 662$\pm$12 & \\ \hline
$m_{\tilde q_L}\pm
\sigma (m_{\tilde q_L})$ & & & & 685$\pm$20& 686$\pm$12 \\ \hline
$m_{\tilde t_1}\pm
\sigma (m_{\tilde t_1})$ & & & & 504$\pm$20& \\ \hline
$m_{\tilde b_1}\pm
\sigma (m_{\tilde b_1})$ & & 276.6$\pm$3.0 & 277.5$\pm$3.0$^{3)}$ & & \\ \hline
$m_{\tilde l_R}\pm
\sigma (m_{\tilde l_R})$ & & & & 156.8$\pm$1.8 & \\ \hline
$m_{\tilde \chi^{\pm}_1}\pm
\sigma (m_{\tilde \chi^{\pm}_1})$& 328.2$\pm$6.5 & & & 232.2$\pm$4.5 & \\ \hline
$m_{\tilde \chi^0_2}\pm
\sigma (m_{\tilde \chi^0_2})$ & 326.2$\pm$6.0 & 96.7$\pm$0.2 & & 230.7$\pm$3.9 & 228.2$\pm$5.0 \\ \hline
$m_{\tilde \chi^0_1}\pm
\sigma (m_{\tilde \chi^0_1})$ & 169.8$^{+0.2}_{-0.8}$ & 44.8$^{+0.1}_{-0.2}$ & & 122.6$^{+0.4}_{-1.0}$ & \\ \hline
$m_{\tilde h^0}\pm
\sigma (m_{\tilde h^0})$ & 97.1$\pm$1.5 & 69$\pm$1.0$^{4)}$ & 69$\pm$1.0$^{4)}$ & 94.7$\pm$1.5 & 94.3$\pm$1.5 \\ \hline
$M^{end\ \ 1)}_{OSSF}
\pm \sigma (M^{end}_{OSSF})$
& & & & & 111.9$\pm$2.5 \\ \hline
$m^{end\ \ 2)}_{OSSF}
\pm \sigma (m^{end}_{OSSF})$
& & 53.3$\pm$0.6 & 52.9$^{+0.1}_{-0.3}$ $^{3)}$ & & \\ \hline\hline
\multicolumn{6}{|c|}{ } \\
\multicolumn{6}{|c|}{
$^{1)}$\ $M^{end}_{OSSF} = m_{\tilde \chi^0_2}\sqrt{1-\Bigl(\frac{m_{\tilde l_R}}{m_{\tilde \chi^0_2}}\Bigr)^2}
\sqrt{1-\Bigl(\frac{m_{\tilde \chi^0_1}}{m_{\tilde l_R}}\Bigr)^2}$ } \\
\multicolumn{6}{|c|}{
$^{2)}$\ $m^{end}_{OSSF} = m_{\chi^0_2} - m_{\chi^0_1}$ } \\
\multicolumn{6}{|c|}{
$^{3)}$ Assuming $m_{\tilde \chi^0_2}= 97$ GeV } \\
\multicolumn{6}{|c|}{
$^{4)}$ From other measurements ($h^0 \rightarrow \gamma \ \gamma$) } \\
\multicolumn{6}{|c|}{ } \\ \hline\hline
\end{tabular}
\label{tb:values}
\end{Tabhere}
\vskip 0.5cm
\begin{Tabhere}
\centering
\caption{\small The relative errors on $m_0$, $m_{1/2}$ and tan$\beta$ in SUGRA points 1, 3 and 5 for each $\rlap/R$
coupling considered.}
\vskip 0.5cm
\begin{tabular}{|c|c|c|c|c|c|c|c|} \hline\hline
& \multicolumn{1}{c|}{ } & \multicolumn{4}{c|}{ } & \multicolumn{2}{c|}{ } \\
Relative errors on the & \multicolumn{1}{c|}{Point 1} & \multicolumn{4}{c|}{Point 3} &
\multicolumn{2}{c|}{Point 5} \\
SUGRA parameters &
\multicolumn{1}{c|}{ } & \multicolumn{4}{c|}{ } & \multicolumn{2}{c|}{ } \\ \cline{2-8}
&$\lambda_{122}$ & $\lambda_{122}$ & $\lambda_{123}$ &
$\lambda_{122}^{*)}$ & $\lambda_{123}^{*)}$ &
$\lambda_{122}$ & $\lambda_{123}$ \\ \hline\hline
$\delta m_{0} / m_0$ (\%) & 12 & 4.4 & 7.3 & 4 & 5.8 & 2.9 & 9.7 \\ \hline
$\delta m_{1/2} / m_{1/2}$ (\%) & 0.3 & 0.3 & 0.6 & 0.2 & 0.4 & 0.5 & 1.4 \\ \hline
$\delta tan \beta / tan \beta$ (\%) & 5 & 3.3 & 3.3 & 1.8 & 1.8 & 6 & 6.2 \\ \hline\hline
\multicolumn{8}{|c|}{ } \\
\multicolumn{8}{|c|}{ $^{*)}$ If the measurement of $m_{h^0}$ is taken into account. } \\
\multicolumn{8}{|c|}{ } \\ \hline\hline
\end{tabular}
\label{tb:sugrafitp1}
\end{Tabhere}
\newpage
\section{Conclusions}
We have studied the feasibility to detect a SUSY signal
by ATLAS in the framework of the SUGRA model
and to determine its parameters in the
case when $R$ parity is broken in conjunction with
lepton number violation: $\lambda \ne 0$. For this purpose we
have chosen three representativ points in the SUGRA
parameter space and two different type of couplings,
both having a value $10^{-3}$, small enough to
concentrate the effect in the LSP decay but large enough
not to see displaced vertex in this decay.
Our conclusions are the following:
1. The SUGRA signal is visible in a very large domain
of the parameter space, even beyond $m_0 \sim m_{1/2} \sim 1$ TeV.
2. The energy scale of the SUGRA signal can be determined by
inclusive measurements like the effective mass ($m_{eff}$) or
the normalised transverse momentum per lepton ($p_t^{l, norm}$).
3. In the case of couplings with absence of a $\tau$ among the
decay products of the $\tilde \chi^0_1$ (e.g. $\lambda_{122}$) one can reconstruct the
SUSY particles and this reconstruction
can be used for a precision determination of the model parameters.
The achieved precision turns out to be
better than it was the case with conserved $R$ parity. This
is because one can reconstruct the LSP from its decay products.
At the low energy point where
the chargino or second lightest neutralino produces additional
leptons this determination is slightly handicapped by the combinatorial
background and the most complex structure.
4. In the case of LSP decay with a $\tau$ particle in the final
state the full reconstruction of the LSP, i.e. the determination of
its four momentum, is not always possible, however,
one can still estimate its mass (except at point 3 - $\lambda_{123}$).
It allows ones to determine the parameters of the SUGRA model in spite of
the large combinatorial background due to the leptonic
decay of the LSP. This determination in most of the
cases is better or at least comparable in precision with that when $R$ parity
is conserved.
\vskip 1.0cm
{\Large \bf Aknowledgements}\\
\\
We would like to thank Daniel Froidevaux for suggesting us to
carry out this study, his constant help and encouragement.
We had many useful discussions within the ATLAS SUSY working
group and would like to express our gratitude to its members, especially
to Ian Hinchliffe, Frank Paige and Giacomo Polesello.
\vskip 1.0cm
|
\section{Experimental Details}\label{appsec:exps}
\subsection{Functional confounders in \gls{gwas}}
Here, we show how $h(\mb{t})= At$ and $A$ reflect the traditional \gls{pca} based adjustment in \gls{gwas}.
Recall population structure acts as a confounder in \gls{gwas}.
\citet{price2006principal} demonstrated that using the principal components of the normalized genetic relationships matrix adjusts for confounding due to population structure in \gls{gwas}.
Let the genotype matrix be $G$ with people as rows and \glspl{snp}{} as columns, such that each element is one of $0,\nicefrac{1}{2}, 1$, where $\nicefrac{1}{2}$ and $1$ refer to one and two copies of the allele respectively at the position of the \gls{snp}.
With $p_s$ as the allele frequency at \gls{snp} $s$~\citep{thorntonsummer}, $\Phi$ is the genetic relationship matrix whose elements are defined as
$\Phi_{i,j} =\frac{1}{S} \sum_{s=1}^S \nicefrac{(G_{i,s} - p_s)(G_{j,s} - p_s)}{p_s(1 - p_s)}$.
Then,~\citet{price2006principal} compute the top $K$ ($10$ suggested) principal components of $\Phi$ to use as the axes of variation due to the population structure.
The eigenvectors of $\Phi$ are the left eigenvectors of $\hat{G}$ such that $\Phi = \hat{G}\hat{G}^T$ which capture independent axes of variation of individuals.
\citet{price2006principal} exploit the idea that if a \gls{snp} aligns with some of the axes of variation, this is due to the population structure.
These axes of variation are the top $K$ eigenvectors $U$ of $\phi = \hat{G}\hat{G}^T \approx U\Lambda U^\top$, where $U\in \mathbb{R}^{N\times K}$, $\Phi\in\mathbb{R}^{N\times N}$ and $\Lambda \in \mathbb{R}^{K\times K}$.
Here, $U$ are also the left singular vectors of $\hat{G} \approx U\Sigma V^T$ where $\Sigma\in \mathbb{R}^{K\times K}$ is diagonal, and $V\in \mathbb{R}^{S\times K}$.
We use $\approx$ to denote that the chosen $K$ eigenvectors explain the variation due to population structure; what remains are random mutations.
Let the $s$th \gls{snp} be $\hat{G}_{\cdot, s}\in \mathbb{R}^{N}$, which is a column in $\hat{G}$.
In \citet{price2006principal}, population structure in the $s$th \gls{snp} is captured in $\hat{G}_{\cdot, s}^\top U$.
In words, projecting the \gls{snp} $\hat{G}_{\cdot, s}$ onto the axes of variation in individuals gives the population structure between $s$th \gls{snp} and the outcome.
This projection $\hat{G}_{\cdot, s}^\top U$ is a row of $\hat{G}^\top U\in \mathbb{R}^{S\times K}$.
In turn, $\hat{G}^\top U\in \mathbb{R}^{S\times K}$ is the population structure in all \glspl{snp}.
Projecting this population structure onto the genotype of an individual gives the confounding due to population structure amongst the \glspl{snp} present in the genotype.
With $G_{j, \cdot}\in \{0,\nicefrac{1}{2}, 1\}^{S}$ as the genotype for an individual $j$, this projection is $\left((\hat{G}^\top U)^\top G_{j, \cdot}\right)$.
However, $\hat{G} \approx U\Sigma V^T$ implies that $\hat{G}^\top U\approx V\Sigma$.
Reflecting this, $h(\mb{t}) = \Sigma V^T \mb{t}$ is the functional confounder for an individual $\mb{t}$.
\newpage
\subsection{Expanded results}\label{appsec:real-exp}
In~\cref{apptab:snps-est}, we list the $13$ \glspl{snp} recovered by \gls{ours}, that have been previously reported as relevant to Celiac disease. In~\cref{fig:tp-fn-rate}, we plot the true positive and false negative rate amongst \glspl{snp} deemed relevant by \gls{ours}.
The ground truth here are the \glspl{snp} reported associated with celiac disease in prior literature. \hspace*{\fill}
\begin{wraptable}{r}{0.45\textwidth}
\centering
\begin{tabular}{ c c c }
\toprule
\gls{snp} & \textsc{Effect} & \textsc{Lasso Coef.} \\
\midrule
rs3748816 & $0.12$ & $0.20$ \\
rs10903122 & $0.10$ & $0.17$ \\
rs2816316 & $0.11$ & $0.20$ \\
rs13151961 & $0.17$ & $0.32$ \\
rs2237236 & $0.17$ & $0.00$ \\
rs12928822 & $0.14$ & $0.29$ \\
rs2187668 & $-0.70$ & $-2.37$ \\
rs2327832 & $-0.12$ & $-0.20$ \\
rs1738074 & $-0.16$ & $-0.23$ \\
rs11221332 & $-0.15$ & $-0.24$ \\
rs653178 & $-0.13$ & $-0.21$ \\
rs4899260 & $-0.12$ & $-0.19$ \\
rs17810546 & $-0.12$ & $-0.20$ \\
\bottomrule \\
\end{tabular}
\caption{Full list of \glspl{snp} previously reported as relevant that were recovered by \gls{ours}, and their
estimated effects and Lasso coefficients for \glspl{snp}. The effect threshold here is $0.1$.}
\label{apptab:snps-est}
\end{wraptable}
\begin{wrapfigure}[15]{r}{0.45\textwidth}
\centering
\includegraphics[width=0.5\textwidth]{images/tp-fn-snps}
\caption{
True positive vs. False negative rate as we vary the threshold on average effects, that determines which \glspl{snp}{} \gls{ours} deems relevant to the outcome.
}
\label{fig:tp-fn-rate}
\end{wrapfigure}
\section{Theoretical details}\label{appsec:theory}
\setcounter{theorem}{0}
\subsection{A note about the assumptions}\label{appsec:assumptions}
\paragraph{Note about the assumptions}
In~\cref{thm:thm-c-red}, assumption 1 consists of three parts that can all be validated on observed data: 1) that the gradient flow converges,
2) that the confounder value of the surrogate matches the confounder value
whose effect is of interest, and
3) that the surrogate intervention lies in the support of the pre-outcome variables.
Assumption 2 is required for expectations and their gradients to exist and be finite.
In~\cref{thm:main-effect-bound}, assumption 1 requires a consistent estimator of $\E[\mb{y} \,\vert\, \mb{t}]$, which can be provided with regression.
Assumption 3 lists regularity conditions which help control how the surrogate estimation error propagates to the effect error.
\subsection{Proof of \Cref{thm:thm-c-red}}\label{appsec:thm-c-red}
We restate the theorem for completeness:
\begin{theorem}
\lemmaone{}
\end{theorem}
\begin{proof}
Recall definition of conditional effect $\phi({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}}_2)) = \E_\mb{\eta} f({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}}_2), \mb{\eta})$.
Recall $\nabla_{\tilde{\mathtt{t}}}$ is the gradient with respect to the first argument of $f$, that is ${\tilde{\mathtt{t}}}$.
First, by assumption 2, $\E$ and $\nabla$ commute, under the dominated convergence theorem. Then, by \gls{c-red}
\[\nabla_{{\tilde{\mathtt{t}}}} \phi({\tilde{\mathtt{t}}}, h(\mathtt{t}^*))^T \nabla_{{\tilde{\mathtt{t}}}} h({\tilde{\mathtt{t}}}) = \nabla_{{\tilde{\mathtt{t}}}} \E_{\mb{\eta}} f({\tilde{\mathtt{t}}},h(\mathtt{t}^*),\mb{\eta})^T \nabla_{{\tilde{\mathtt{t}}}} h({\tilde{\mathtt{t}}}) = \E_{\mb{\eta}} [\nabla_{{\tilde{\mathtt{t}}}} f({\tilde{\mathtt{t}}},h(\mathtt{t}^*),\mb{\eta})^T \nabla_{{\tilde{\mathtt{t}}}} h({\tilde{\mathtt{t}}})] = 0 .\]
Now consider the gradient flow equation $\nicefrac{d{\tilde{\mathtt{t}}}(s)}{ds} = -\nabla_{{\tilde{\mathtt{t}}}} (h({\tilde{\mathtt{t}}}) - h(\mathtt{t}^*_2))^2$.
We refer to the gradient evaluated at ${\tilde{\mathtt{t}}}$ as $\Delta{\tilde{\mathtt{t}}} = - \nabla_{{\tilde{\mathtt{t}}}} (h({\tilde{\mathtt{t}}}) - h(\mathtt{t}^*_2))^2 = - 2 (h({\tilde{\mathtt{t}}}) - h(\mathtt{t}^*_2))\nabla_{{\tilde{\mathtt{t}}}} h({\tilde{\mathtt{t}}})$.
We will express $\phi(\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2)), h(\mathtt{t}^*_2))$ as defined by the starting point $\phi(\mathtt{t}^*, h(\mathtt{t}^*_2))$ and the gradient flow equation.
Let the solution path to the gradient flow equation be $C$ with $\mathtt{t}^*, \mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))$ being the starting and ending points respectively.
By the Gradient Theorem~\citep{spivak2018calculus}, we have that $\phi(\mathtt{t}^*, h(\mathtt{t}^*_2))$ and $\phi(\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2)), h(\mathtt{t}^*_2))$ are related via the line integral over $C$:
\[\int_C \nabla_{{\tilde{\mathtt{t}}}} \phi({\tilde{\mathtt{t}}}, h(\mathtt{t}^*_2)) \cdot d {\tilde{\mathtt{t}}} = \phi(\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2)), h(\mathtt{t}^*_2)) - \phi({\tilde{\mathtt{t}}}, h(\mathtt{t}^*_2)) \]
Let ${\tilde{\mathtt{t}}}(s)$ be a parametrization of solution path $C$ by the scalar time $s\in[0,\infty)$.
Now, to obtain the value of $\phi({\tilde{\mathtt{t}}}, h(\mathtt{t}^*_2))$, we will compute the line integral over the vector field defined by $\nabla_{{\tilde{\mathtt{t}}}} \phi({\tilde{\mathtt{t}}}, h(\mathtt{t}^*_2))$, which exists by assumption 2 in~\cref{thm:thm-c-red}, evaluated along the path $C$ defined by $\Delta {\tilde{\mathtt{t}}}(s)$:
\begin{align}\label{eq:f-function-change}
\begin{split}
\phi(\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2)), h(\mathtt{t}^*_2)) &= \phi(\mathtt{t}^*, h(\mathtt{t}^*_2)) + \int_C \nabla_{{\tilde{\mathtt{t}}}} \phi({\tilde{\mathtt{t}}}, h(\mathtt{t}^*_2)) \cdot d {\tilde{\mathtt{t}}}
\\
&= \phi(\mathtt{t}^*, h(\mathtt{t}^*_2)) + \int_0^\infty \nabla_{{\tilde{\mathtt{t}}}} \phi({\tilde{\mathtt{t}}}(s), h(\mathtt{t}^*_2))^T \fullder{{\tilde{\mathtt{t}}}(s)}{s}\ ds
\\
& = \phi(\mathtt{t}^*, h(\mathtt{t}^*_2)) + \int_0^\infty \nabla_{{\tilde{\mathtt{t}}}} \phi({\tilde{\mathtt{t}}}(s), h(\mathtt{t}^*_2))^T \Delta {\tilde{\mathtt{t}}}(s)\ ds
\\
& = \phi(\mathtt{t}^*, h(\mathtt{t}^*_2))
\\ & \quad \quad + \int_0^\infty - 2((h({\tilde{\mathtt{t}}}(s)) - h(\mathtt{t}^*_2)))\, \nabla_{{\tilde{\mathtt{t}}}} \phi({\tilde{\mathtt{t}}}(s), h(\mathtt{t}^*_2))^T \nabla_{{\tilde{\mathtt{t}}}} h({\tilde{\mathtt{t}}}(s))\ ds
\\
& = \phi(\mathtt{t}^*, h(\mathtt{t}^*_2)) + 0 \quad \quad \text{\{by \gls{c-red}\}}
\end{split}
\end{align}
Finally, by assumption 1 in~\cref{thm:thm-c-red}, $h(\mathtt{t}^\prime(\mathtt{t}^*,h(\mathtt{t}^*_2))) = h(\mathtt{t}^*_2)$, and so
\begin{align}\label{eq:f-connection}
\phi(\mathtt{t}^*, h(\mathtt{t}^*_2)) = \phi(\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2)), h(\mathtt{t}^*_2))
= \phi(\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2)), h(\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))))
\end{align}
For clarity, the same equation, but using $\mathtt{t}^\prime$ and suppressing dependence on $\mathtt{t}^*, h(\mathtt{t}^*_2)$):
\begin{align}\label{eq:f-connection}
\phi(\mathtt{t}^*, h(\mathtt{t}^*_2)) = \phi(\mathtt{t}^\prime, h(\mathtt{t}^*_2))
= \phi(\mathtt{t}^\prime, h(\mathtt{t}^\prime))
\end{align}
Under the causal model for \gls{efc}, the outcome $\mb{y} = f(\mb{t}, h(\mb{t}), \mb{\eta})$.
Then, $\forall {\tilde{\mathtt{t}}} \in \textrm{supp}(p(\mb{t}))$,
\begin{align}
\label{eq:equality-of-conditional}
\quad \E[\mb{y}\,\vert\, \mb{t} = {\tilde{\mathtt{t}}}] = \E_{\mb{\eta}}[f({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}}), \mb{\eta})] = \phi({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}})).
\end{align}
Using that $ \mathtt{t}^\prime(\mathtt{t}^*, \mathtt{t}^*_2) \in \textrm{supp}(p(\mb{t}))$ and~\cref{eq:equality-of-conditional,eq:f-connection}, the conditional effect is identified
\begin{align}
\begin{split}
\phi(\mathtt{t}^*, h(\mathtt{t}^*_2)) & = \phi(\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2)), h(\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))))
\\
& = \E[\mb{y}\,\vert\, \mb{t}=\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))]
\end{split}
\end{align}
Thus, the conditional effect, and consequently the average effect{}, are identified as $\E[\mb{y}\,\vert\, \mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))]$ and ${\tau}(\mathtt{t}^*) = \E_{h(\mb{t})}\E[\mb{y}\,\vert\, \mathtt{t}^\prime(\mathtt{t}^*, h(\mb{t}))]$ respectively.
\end{proof}
\paragraph{Note about convergence of gradient flow}
Any ODE's solution, if it exists and converges, converges to an $\omega$-limit set~\cite{teschl2012ordinary}.
An $\omega$-limit set is nonempty when the solution path lies entirely in a closed and bounded set and can consist of limit cycles, equilibrium points, or neither~\cite{hirsch1974differential,teschl2012ordinary}.
A gradient flow equation $\nicefrac{d {\tilde{\mathtt{t}}} (s)}{ ds} = -\nabla h({\tilde{\mathtt{t}}})$ (also called a gradient system) has the special property that its $\omega$-limit set only consists of critical points of $h({\tilde{\mathtt{t}}})$; critical points of $h({\tilde{\mathtt{t}}})$ are also equilibrium points of the gradient flow equation~\cite{hirsch1974differential}.
Further, if $\nabla h({\tilde{\mathtt{t}}})$ exists and is bounded and $h({\tilde{\mathtt{t}}})$ has bounded sublevel sets ($\{{\tilde{\mathtt{t}}}: h({\tilde{\mathtt{t}}}) \leq c\}$), then the solution to the gradient flow equation will entirely lie within a bounded set.
This is because along the solution path, $h({\tilde{\mathtt{t}}}(s))$ always decreases meaning that the solution will remain in any sublevel set it started in.
Thus, if $h({\tilde{\mathtt{t}}})$ has bounded sublevel sets, the solution of the gradient flow equation will converge only to critical points of $h({\tilde{\mathtt{t}}})$.
\subsection{Estimation error in \gls{ours}}\label{appsec:lode-error}
\begin{theorem}
\theoremmain
\end{theorem}
\begin{proof} (of \Cref{thm:main-effect-bound})
Recall the definition of conditional effect : $\phi({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}}_2)) = \E_{\mb{\eta}}f({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}}_2), \mb{\eta})$.
\gls{ours}'s estimate of the conditional effect is $\hat{f}(\hat{\mathtt{t}}(\mathtt{t}^*, h(\mathtt{t}^*_2)))$.
We will suppress notation for dependence on $\mathtt{t}^*, h(\mathtt{t}^*_2)$, and use $\mathtt{t}^\prime$ and $\hat{\mathtt{t}}$ to refer to the true surrogate intervention and the estimated surrogate interventions respectively.
Note $\hat{f}$ is the estimate of the conditional expectation $\E[\mb{y}\,\vert\, \mb{t}={\tilde{\mathtt{t}}}]$, learned from $N$ samples.
We first bound the error by splitting into two parts and bounding each separately:
\begin{align*}
|\xi(\mathtt{t}^*,h(\mathtt{t}^*_2))| & = | \hat{f}(\hat{\mathtt{t}}) - \phi(\mathtt{t}^*, h(\mathtt{t}^*_2)) |
\\
&\leq| \hat{f}(\hat{\mathtt{t}}) - \phi(\hat{\mathtt{t}}, h(\hat{\mathtt{t}})) | + | \phi(\hat{\mathtt{t}}, h(\hat{\mathtt{t}})) - \phi(\mathtt{t}^*, h(\mathtt{t}^*_2))|
\\
& \leq c(N) + | \phi(\hat{\mathtt{t}}, h(\hat{\mathtt{t}})) - \phi(\mathtt{t}^*, h(\mathtt{t}^*_2))|
\\
& \leq |\phi(\hat{\mathtt{t}}, h(\hat{\mathtt{t}})) - \phi(\hat{\mathtt{t}}, h(\mathtt{t}^*_2)) | + |\phi(\hat{\mathtt{t}}, h(\mathtt{t}^*_2)) - \phi(\mathtt{t}^*, h(\mathtt{t}^*_2)) | + c(N)
\end{align*}
The first term is bounded via the Lipschitz-ness of $\phi$ as a function of $h({\tilde{\mathtt{t}}})$ with fixed first argument ${\tilde{\mathtt{t}}} = \hat{\mathtt{t}}$.
\[|\phi(\hat{\mathtt{t}}, h(\hat{\mathtt{t}})) - \phi(\hat{\mathtt{t}}, h(\mathtt{t}^*_2)) | \leq L_{z,\hat{\mathtt{t}}} |h(\hat{\mathtt{t}}) - h(\mathtt{t}^*_2)|\]
We now bound the remaining term.
Recall that \Gls{ours}'s computation of the surrogate intervention involved $K$ gradient steps, each of size $\ell$.
We work with a constant step-size but the analysis can be generalized to a non-uniform step size.
Indexing steps with $i$, let $d_i = h({\tilde{\mathtt{t}}}_i) - h(\mathtt{t}^*_2)$ be the confounder mismatch error at the $i$th iterate.
Then note that $\hat{\mathtt{t}} = \mathtt{t}^* - \ell \sum_{i=0}^{K-1} 2 d_i \nabla_{\tilde{\mathtt{t}}} h({\tilde{\mathtt{t}}}_i) $.
We can use this to bound the error $\phi(\hat{\mathtt{t}}, h(\mathtt{t}^*_2)) - \phi(\mathtt{t}^*, h(\mathtt{t}^*_2)) $.
With ${\tilde{\mathtt{t}}}_K = \hat{\mathtt{t}}$ and ${\tilde{\mathtt{t}}}_0 = \mathtt{t}^*$, we proceed by expressing the error as a telescoping sum and using the Taylor expansion for $\phi({\tilde{\mathtt{t}}}, h(\mathtt{t}^*_2))$ in terms of the the first argument ${\tilde{\mathtt{t}}}$.
\allowdisplaybreaks
\begin{align}\label{eq:lode-est-err}
&\phi(\hat{\mathtt{t}}, h(\mathtt{t}^*_2)) - \phi(\mathtt{t}^*, h(\mathtt{t}^*_2)) = \sum_{i=0}^{K-1} \phi({\tilde{\mathtt{t}}}_{i+1}, h(\mathtt{t}^*_2)) - \phi({\tilde{\mathtt{t}}}_{i}, h(\mathtt{t}^*_2))
\\
& = \sum_{i=0}^{K-1} \nabla_{\tilde{\mathtt{t}}} \phi({\tilde{\mathtt{t}}}_i, h(\mathtt{t}^*_2))^\top ({\tilde{\mathtt{t}}}_{i+1} - {\tilde{\mathtt{t}}}_i)
\\ & \quad\quad \quad \quad +
\frac{1}{2}({\tilde{\mathtt{t}}}_{i+1} - {\tilde{\mathtt{t}}}_i)^\top \nabla_{\tilde{\mathtt{t}}}^2 \phi({\tilde{\mathtt{t}}}_{i}, h(\mathtt{t}^*_2)) ({\tilde{\mathtt{t}}}_{i+1} - {\tilde{\mathtt{t}}}_i) + \mathcal{O}(\|{\tilde{\mathtt{t}}}_{i+1} - {\tilde{\mathtt{t}}}_i\|_2^3)
\\
& = \sum_{i=0}^{K-1} 2 \ell d_i \nabla_{\tilde{\mathtt{t}}} \phi({\tilde{\mathtt{t}}}_i, h(\mathtt{t}^*_2))^\top \nabla_{\tilde{\mathtt{t}}} h({\tilde{\mathtt{t}}}_i)
+ 2(\ell d_i)^2 \nabla_{\tilde{\mathtt{t}}} h({\tilde{\mathtt{t}}}_i)^\top \nabla_{\tilde{\mathtt{t}}}^2 \phi({\tilde{\mathtt{t}}}_{i}, h(\mathtt{t}^*_2)) \nabla_{\tilde{\mathtt{t}}} h({\tilde{\mathtt{t}}}_i)+ \mathcal{O}(\ell^3)
\\
& = \sum_{i=0}^{K-1} 0 + 2(\ell d_i)^2 \nabla_{\tilde{\mathtt{t}}} h({\tilde{\mathtt{t}}}_i)^\top \nabla_{\tilde{\mathtt{t}}}^2 \phi({\tilde{\mathtt{t}}}_{i}, h(\mathtt{t}^*_2)) \nabla_{\tilde{\mathtt{t}}} h({\tilde{\mathtt{t}}}_i) + \mathcal{O}(\ell^3)
\\
& = \mathcal{O}(K\ell^3) + \sum_{i=0}^{K-1} 2(\ell d_i)^2 \nabla_{\tilde{\mathtt{t}}} h({\tilde{\mathtt{t}}}_i)^\top \nabla_{\tilde{\mathtt{t}}}^2 \phi({\tilde{\mathtt{t}}}_{i}, h(\mathtt{t}^*_2)) \nabla_{\tilde{\mathtt{t}}} h({\tilde{\mathtt{t}}}_i)
\\
& \leq \mathcal{O}(K\ell^3) + \sum_{i=0}^{K-1} 2(\ell (h({\tilde{\mathtt{t}}}_i) - h(\mathtt{t}^*_2)))^2 \left|\nabla_{\tilde{\mathtt{t}}} h({\tilde{\mathtt{t}}}_i)^\top \nabla_{\tilde{\mathtt{t}}}^2 \phi({\tilde{\mathtt{t}}}_{i}, h(\mathtt{t}^*_2)) \nabla_{\tilde{\mathtt{t}}} h({\tilde{\mathtt{t}}}_i) \right|
\\
& \leq \mathcal{O}(K\ell^3) + \sum_{i=0}^{K-1}2\ell^2M\left| \nabla_{\tilde{\mathtt{t}}} h({\tilde{\mathtt{t}}}_i)^\top \nabla_{\tilde{\mathtt{t}}}^2 \phi({\tilde{\mathtt{t}}}_{i}, h(\mathtt{t}^*_2)) \nabla_{\tilde{\mathtt{t}}} h({\tilde{\mathtt{t}}}_i) \right|
\\
& \leq \mathcal{O}(K\ell^3) + \sum_{i=0}^{K-1} 2\ell^2M \sigma_{\mathtt{H} \phi } \|\nabla_{\tilde{\mathtt{t}}} h({\tilde{\mathtt{t}}}_i)\|_2^2
\\
& \leq \mathcal{O}(K\ell^3) + \sum_{i=0}^{K-1} 2\ell^2M \sigma_{\mathtt{H} \phi } L^2_{h}
\\
& = 2K \ell^2 \left(\mathcal{O}(\ell) + M \sigma_{\mathtt{H} \phi } L^2_{h} \right),
\end{align}
where the inequalities follow by the maximum value of $(h({\tilde{\mathtt{t}}}_i) - h(\mathtt{t}^*_2))^2$, bounded eigenvalues of the Hessian of $\phi$ and the Lipschitz-ness of $h({\tilde{\mathtt{t}}})$.
Another way we bound the error is via the Lipschitz constant of the conditional expectation as a function of ${\tilde{\mathtt{t}}}$.
Recall this is $L_{e}$. An alternate bound on the error is as follows:
\begin{align*}
|\phi(\hat{\mathtt{t}}, h(\hat{\mathtt{t}})) - \phi(\mathtt{t}^*, h(\mathtt{t}^*_2))| = |\phi(\hat{\mathtt{t}}, h(\hat{\mathtt{t}})) - \phi(\mathtt{t}^\prime, h( \mathtt{t}^\prime))| \leq & L_e\|\mathtt{t}^\prime - \hat{\mathtt{t}}\|_2
\end{align*}
The bound follows:
\begin{align*}
|\xi({\tilde{\mathtt{t}}},h(\mathtt{t}^*_2))| & \leq c(N) + \min\left(L_e\|\mathtt{t}^\prime - \hat{\mathtt{t}}\|_2,
\quad 2K \ell^2 \left(\mathcal{O}(\ell) + M \sigma_{\mathtt{H} \phi } L^2_{h} \right) +
L_{z,\hat{\mathtt{t}}} \|h(\hat{\mathtt{t}}) - h(\mathtt{t}^*_2)\|_2\right)
\end{align*}
\end{proof}
\subsubsection{A note on linear confounder functions and \gls{ours}}\label{appsec:linear-conf-func}
In the proof above, the error in Euler integration accumulates due to terms like this one: $\nabla_{\tilde{\mathtt{t}}}^\top h({\tilde{\mathtt{t}}}) \nabla^2_{\tilde{\mathtt{t}}} f({\tilde{\mathtt{t}}}, h(\mathtt{t}^*),\eta) \nabla_{\tilde{\mathtt{t}}} h({\tilde{\mathtt{t}}})$.
For a linear confounder function that satisfies $\nabla_{\tilde{\mathtt{t}}} h({\tilde{\mathtt{t}}}) = \beta$, such terms can be expressed as $\beta^\top \nabla_{\tilde{\mathtt{t}}} (\nabla_{\tilde{\mathtt{t}}} f({\tilde{\mathtt{t}}}, h(\mathtt{t}^*),\eta)^\top \beta) = \beta^\top \nabla_{\tilde{\mathtt{t}}} (0)= 0 $ under \gls{c-red}.
Thus, such error does not accumulate even with large step sizes.
Further, note that the gradient flow equation in \gls{ours} for the causal model $A$ in~\cref{sec:exps} is a linear ODE whose solution has a closed form expression and one can estimate the surrogate without numerical integration~\cite{teschl2012ordinary}.
\subsection{Proof of sufficiency of Effect Connectivity}\label{appsubsec:sufficiency}
\begin{theorem}
\sufficiencylemma{}{}
\end{theorem}
\begin{proof}
Recall $\phi({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}})) = \E_\mb{\eta} f({\tilde{\mathtt{t}}},h({\tilde{\mathtt{t}}}), \mb{\eta})$.
We have $\forall \mathtt{t}^* \in\textrm{supp}(p(\mb{t}))$:
\[ p(h(\mb{t}) = h(\mathtt{t}^*_2))>0\implies p(\phi(\mb{t}, h(\mb{t})) = \phi(\mathtt{t}^*, h(\mathtt{t}^*_2))\,\vert\, h(\mb{t}) = h(\mathtt{t}^*_2)) >0.\]
This implies $\exists \mathtt{t}^\prime\in \textrm{supp}(\mb{t}), \phi(\mathtt{t}^\prime, h(\mathtt{t}^*_2)) = \phi(\mathtt{t}^*, h(\mathtt{t}^*_2)), \quad s.t. \quad h(\mathtt{t}^\prime) = h(\mathtt{t}^*_2).$
Then, $\phi(\mathtt{t}^*, h(\mathtt{t}^*_2)) = \phi(\mathtt{t}^\prime, h(\mathtt{t}^*_2)) = \phi(\mathtt{t}^\prime, h(\mathtt{t}^\prime)) = \E[\mb{y}\,\vert\, \mb{t}=\mathtt{t}^\prime].$
\end{proof}
\subsection{Necessity of Effect Connectivity for Nonparametric effect estimation in \gls{efc}}\label{appsubsec:necessity-proof}
\begin{theorem}
\necessitylemma{}
\end{theorem}
\begin{proof} (Proof of~\Cref{thm:thm-necessity})
Let the outcome be $\mb{y}= f(\mb{t}, h(\mb{t}))$.
Recall the joint distribution $p(\mb{t}, \mb{y})$ and let $h(\mb{t})$ be the confounder.
Let Effect Connectivity be violated, i.e.\ there exists a non-measure-zero subset $B \in \textrm{supp}(\mb{t})\times \textrm{supp}(h(\mb{t}))$ such that \footnote{Non-zero w.r.t.\ the product measure over $\textrm{supp}(\mb{t})\times \textrm{supp}(h(\mb{t}))$ due to $p$.}:
\[\forall \,\, {\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}}_2) \in B, \quad \quad p(f(\mb{t}, h(\mb{t})) = f({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}}_2))\,\vert\, h(\mb{t}) = h({\tilde{\mathtt{t}}}_2)) = 0.\]
Now, we construct a new outcome $\mb{y}_2 = f_2(\mb{t}, h(\mb{t}))$ and show the conditional effects for this new outcome are different from the one defined by $f$ on $\forall ({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}}_2))\in B$. Let
\[f_2({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}}_2)) = f({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}}_2)) + 10*1(({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}}_2))\in B)|.\]
We have $f_2({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}})) = f({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}})) \, \forall {\tilde{\mathtt{t}}} \in \textrm{supp}(\mb{t})$ , as the additional term in $f_2$ is only present for $({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}}_2))\in B$; this follows from the fact that $\forall {\tilde{\mathtt{t}}} \in \textrm{supp}(\mb{t})$, $({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}})) \not\in B$ as
\[p[f(\mb{t}, h(\mb{t})) = f({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}})) \,\vert\, h(\mb{t}) = h({\tilde{\mathtt{t}}})] = p[f(\mb{t}, h(\mb{t})) = f({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}}))] >0.\]
Thus, $p(\mb{y}, \mb{t}) =^d p(\mb{y}_2, \mb{t})$ are equal in distribution since $B \cap \textrm{supp}(\mb{t}, h(\mb{t})) = \emptyset$.
This means that the conditional effects are different for the outcomes $\mb{y}, \mb{y}_2$ for all $({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}}_2)) \in B$:
\[\E[\mb{y} \,\vert\, do(\mb{t} = {\tilde{\mathtt{t}}}), h(\mb{t}) = h({\tilde{\mathtt{t}}}_2)] \not= \E[\mb{y}_2 \,\vert\, do(\mb{t} = {\tilde{\mathtt{t}}}), h(\mb{t}) = h({\tilde{\mathtt{t}}}_2)] \]
Therefore, for causal models that violates Effect Connectivity, there exist observationally equivalent causal models with different causal effects.
Thus, nonparametric effect estimation is impossible.
Thus, Effect Connectivity is required for \gls{efc}.
\end{proof}
\subsection{Algorithmic details}
We give in ~\cref{alg:full-eff-estimation} pseudocode for \gls{ours}.
\begin{algorithm}[ht]
\SetAlgoLined
\SetKwProg{Fn}{Function}{ is}{end}
\KwIn{Functional confounder $h(\mb{t})$; tolerance $\epsilon$}
\KwOut{Conditional effects of $\mathtt{t}^*, h(\mathtt{t}^*_2)$}
Regress $\mb{y}$ on $\mb{t}$ and compute
$\hat{f}() := \arg\min_{u\in \mathcal{F}} \E_{\mb{y},\mb{t}}(\mb{y} - u(\mb{t}))^2 .$ \\
To estimate effects of $\mathtt{t}^*, h(\mathtt{t}^*_2)$, compute the surrogate intervention $\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))$ by Euler integrating the gradient flow equation, initialized at ${\tilde{\mathtt{t}}} = \mathtt{t}^*$, until $(h({\tilde{\mathtt{t}}}_s) - h(\mathtt{t}^*_2))^2 < \epsilon$.
\[\fullder{{\tilde{\mathtt{t}}}(s)}{s} = \nabla_{\tilde{\mathtt{t}}} (h({\tilde{\mathtt{t}}}_s) - h(\mathtt{t}^*_2))^2,\]
\\
Return $\hat{f}(\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2)))$;
\caption{\gls{ours} for $do(\mb{t}=\mathtt{t}^*)$}
\label{alg:full-eff-estimation}
\end{algorithm}
\paragraph{Extensions of \gls{ours}}
Consider that we have access to $m(h(\mb{t}))$ for some bijective differentiable function $m(\cdot)$, instead of $h(\mb{t})$.
The orthogonality in \gls{c-red} holds $\nabla_{{\tilde{\mathtt{t}}}} f({\tilde{\mathtt{t}}},h({\tilde{\mathtt{t}}}_2), \eta)^T \nabla_{{\tilde{\mathtt{t}}}} m(h({\tilde{\mathtt{t}}})) = m'(h({\tilde{\mathtt{t}}}))\nabla_{{\tilde{\mathtt{t}}}} f({\tilde{\mathtt{t}}},h({\tilde{\mathtt{t}}}_2), \eta)^T \nabla_{{\tilde{\mathtt{t}}}} h({\tilde{\mathtt{t}}}) = 0$.
Then, using $m(h({\tilde{\mathtt{t}}}))$ to compute the surrogate $\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))$, \gls{ours} would estimate valid effects.
Similarly, \gls{ours} can estimate the effect on any differentiable transformation of the outcome $m(y)$, because
$\nabla_{{\tilde{\mathtt{t}}}} m(y_{\tilde{\mathtt{t}}})^T\nabla_{{\tilde{\mathtt{t}}}} h({\tilde{\mathtt{t}}}) = m'(y_{\tilde{\mathtt{t}}})\nabla_{{\tilde{\mathtt{t}}}} f({\tilde{\mathtt{t}}},h({\tilde{\mathtt{t}}}_2), \eta)^T \nabla_{{\tilde{\mathtt{t}}}} h({\tilde{\mathtt{t}}}) = 0$ holds.
\section{Introduction}
Determining the effect of interventions on outcomes using observational data lies at the core of many fields like medicine, economic policy, and genomics.
For example, policy makers estimate effects to elect whether to invest in education or job training programs.
In medicine, doctors use effects to design optimal treatment strategies for patients.
Geneticists perform \gls{gwas} to relate genotypes and phenotypes.
In observational data, there could exist unobserved variables that affect both the intervention and the outcome, called confounders.
A necessary condition for the causal effect to be identified is that all confounders are observed; called \emph{ignorability}.
If ignorability holds, a sufficient condition for causal effect estimation is adequate variation in the intervention after conditioning on the confounders; this condition is called \emph{positivity}.
The data apriori does not differentiate between confounders and interventions.
It is the practitioners that select interventions of interest from all pre-outcome variables (variables that occur before the outcome).
Then, assuming knowledge of the data generating mechanism, practitioners can label certain variables amongst the remaining pre-outcome variables as confounders.
This corresponds to indexing into the set of pre-outcome variables.
In certain problems the confounders are specified as
a function of the pre-outcome variables that does not simply index
into the set of pre-outcome variables. For a concrete example, consider
\gls{gwas}. The goal in \gls{gwas} is to estimate the influence of
genetic variations on phenotypes like disease risk.
In \gls{gwas},
population and family structures both result in certain genetic variations and affect phenotypes and therefore, are confounders~\cite{astle2009population}.
Practitioners specify these confounders by using the genetic similarity between individuals~\cite{lippert2011fast,price2006principal,yu2006unified}, which is
a function of the genetic variations.
When the confounders are a function of the same pre-outcome variables that define the interventions, positivity is violated.
Then, the class of interventions whose effects are estimable is not well-defined.
We study causal effect estimation in such settings, where a function of the
pre-outcome variables provides the confounder and these same pre-outcome variables
define the intervention. We call this \glsreset{efc}\gls{efc}.
In \gls{efc}, one column in the observed data is the outcome and all others are pre-outcome variables.
We assume access to a function $h(\cdot)$ that takes as input the pre-outcome variables
and returns the value of the confounder.
Further, we assume these confounders give us ignorability.
In settings like \gls{gwas}, the function $h$ reflects the practitioner-specified function that captures the genetic variation influenced by the population structure.
In traditional \gls{ci}\xspace, $h(\cdot)$ reflects the selection of certain variables in the data and labelling them as confounders.
In \gls{efc}, two different values of the confounder are never observed for the same setting of the pre-outcome variables.
This means that positivity is violated and the effects of only certain interventions may be estimable.
We address this issue in two ways.
First, we investigate a class of plausible interventions that are \emph{functions} of the observed pre-outcome variables, called functional interventions.
We develop a sufficient condition to estimate the effects of said functional interventions, called \glsreset{func-pos}\gls{func-pos}.
Second, we consider intervening on all pre-outcome variables, called the \emph{full} intervention.
We develop a sufficient condition to estimate the effect of the \emph{full} intervention, called \gls{c-red}.
For an intervention, given a confounder value, \gls{c-red} allows us to compute a surrogate intervention such that the conditional effect of the surrogate is equal to that of the original intervention.
We also show that such surrogate interventions exist only under a certain condition that we call Effect Connectivity, that is necessary for nonparametric effect estimation in \gls{efc}.
This condition is satisfied by default in traditional \gls{ci}\xspace if ignorability and positivity hold.
Then, we develop an algorithm for causal estimation assuming \gls{c-red}, called \glsreset{ours}\gls{ours}, which estimates effects using surrogate interventions.
If the surrogate is not estimated well, \gls{ours}'s estimates are biased.
We establish bounds on this bias that capture the mitigating effect of the smoothness of the true outcome function.
\paragraph{Related work}
The problem of \glsreset{gwas}\gls{gwas} is to estimate the effect of genetic variations(also called \glspl{snp}) on the phenotype~\cite{visscher201710}.
The ancestry of the subjects acts as a confounder in \gls{gwas}.
In \gls{gwas} practice, \gls{pca} and \glspl{lmm} are used to compute this confounding structure~\cite{price2006principal,yu2006unified}.
\citet{lippert2011fast} suggest estimating the confounders and effects on \emph{separate} subsets of the \glspl{snp}.
This separation disregards the confounding that is captured in the interaction of the two subsets of \glspl{snp}.
\Gls{gwas} is a special case of \glsreset{mci}\gls{mci} where the confounder value
is specified via optimization as a function of the pre-outcome variables~\cite{ranganath2018multiple,wang2018blessings}.
In all these settings, positivity is violated and not all effects are estimable.
We provide an avenue for nonparametric effect-estimation of the full intervention under a new condition, \gls{c-red}.
\paragraph{\glsreset{ci}Traditional \gls{ci}\xspace review}\label{sec:obs-ci}
We setup causal inference with Structural Causal Models \citep{pearl2009causal} and use $do(\mb{t}=\mathtt{t}^*)$ to denote making an intervention.
Let $\mb{t}$ be a vector of the interventions, $\mb{z}$ be the confounder, and $\mb{y}$ be the outcome.
Let $\mb{\eta} \sim p(\mb{\eta}) (\mb{\eta} \rotatebox[origin=c]{90}{$\models$} (\mb{z}, \mb{t}))$ be noise.
With $f$ as the \emph{outcome function}, we define the causal model for traditional \gls{ci}\xspace as
\footnote{We focus on $f$ that generates $\mb{y}$ from $\mb{t}, \mb{z}$. SCMs generally specify the function that generates $\mb{t}$ from $\mb{z}$ also.}:
\begin{align*}
\mb{z} \sim p(\mb{z}), \quad \mb{t} \sim p(\mb{t} \,\vert\, \mb{z}), \quad y = f(\mb{t}, \mb{z}, \mb{\eta}).
\end{align*}
Let $p(y,\mb{z},\mb{t})$ denote the joint distribution implied by this data generating process.
The effects of interest under the full intervention $do(\mb{t}=\mathtt{t}^*)$ are the average and~\emph{\condeff}
\begin{align}
(\text{average})\quad &\tau(\mathtt{t}^*)=\E_{\mb{z}, \mb{\eta}}f(\mathtt{t}^*, \mb{z}, \mb{\eta})
\quad \quad \quad
(\text{conditional}) \quad \phi(\mathtt{t}^*,{\mathtt{z}})=\E_{\mb{\eta}}\left[f(\mathtt{t}^*, {\mathtt{z}}, \mb{\eta})\right].
\end{align}
With observed confounders, two assumptions make causal estimation possible: \emph{ignorability} and \emph{positivity}.
Ignorability means that \emph{all} confounders $\mb{z}$ are observed in data.
Conditioning on all the confounders, the outcome under an intervention is distributed as if conditional on the value of the intervention:
$p(\mb{y} = y_1 \,\vert\, do(\mb{t}=\mathtt{t}^*), \mb{z}={\mathtt{z}}) = p(f(\mathtt{t}^*, {\mathtt{z}}, \mb{\eta}) = y_1) = p(\mb{y} = y_1 \,\vert\, \mb{t}=\mathtt{t}^*, \mb{z}={\mathtt{z}}) $.
This allows the expression of average effect\ as an expectation over the \emph{observed} outcomes
$\tau(\mathtt{t}^*) = \E_{\mb{z}, \mb{\eta}}[f(\mathtt{t}^*, \mb{z}, \mb{\eta})] = \E_{\mb{z}}\E[y \,\vert\, \mb{z}, \mathtt{t}^*].$
The conditional expectation only exists for all $\mathtt{t}^*$ if $p(y \,\vert\, \mb{z}, \mb{t}=\mathtt{t}^*) = \nicefrac{p(y, \mb{z}, \mb{t}=\mathtt{t}^*)}{p(\mb{z}) p(\mb{t} =\mathtt{t}^* \,\vert\, \mb{z})}$ exists.
\emph{Positivity} guarantees this existence
\begin{align}
(\text{positivity})\quad \forall \mathtt{t}^* \in \textrm{supp}(\mb{t}) \quad p(\mb{z}={\mathtt{z}})>0\implies p(\mb{t} =\mathtt{t}^* \,\vert\, \mb{z} = {\mathtt{z}}) > 0.
\label{eq:positivity}
\end{align}
\section{\Acrlong{efc}}\label{sec:est-without-positivity}
In traditional \gls{ci}\xspace, causal estimation relied on knowing the confounders.
In this section, we consider settings where confounders are known via a function of the pre-outcome variables $h(\mb{t})=\mb{z}$.
We call this setting~\glsreset{efc}\emph{\gls{efc}}.
An example of this is~\gls{gwas}, where \glspl{snp} (the pre-outcome variables) are used to estimate the confounding population structure through methods like \gls{pca}~\citep{yu2006unified}.
Assuming the confounders are a function of the pre-outcome variables violates positivity in general.
Positivity is violated in this setting because
\[\forall t_1, t_2 \in \textrm{supp}(\mb{t}) \,\,\, s.t. \,\,\, h(t_2) \not = h(t_1)\,\, \implies p(\mb{z} = h(t_2) \,\vert\, \mb{t}=t_1) = 0 \not = p(\mb{z} = h(t_2)) > 0\]
In words, two different confounder values cannot occur for the same $t$.
A positivity violation precludes nonparametric effect estimation of the full intervention $do(\mb{t}=\mathtt{t}^*)$.
\paragraph{Positivity and Regression Identifiability}
Positivity can be viewed as providing identifiability.
To see this, let the confounder be $\mb{z}=h(\mb{t})$ and the outcome be
$y(\mb{t}, \mb{z}, \mb{\eta}) = \mb{z} + h(\mb{t}).$
Now consider regressing $\mb{z}$ and $\mb{t}$ onto $y$.
Then, functions $y = \alpha \mb{z} + \beta h(\mb{t})$ indexed by $\alpha,\beta$, such that $\alpha + \beta=2$, are consistent with the observed data.
Thus, there exist infinitely many solutions to the conditional expectation of $y$ on $(\mb{t},\mb{z})$, meaning that the regression is not identifiable.
Assuming positivity necessitates sufficient randomness to identify the regression and thus the causal effect.
A violation of positivity means that nonparametric estimation of causal effects needs further assumptions.
\subsection{Setup for \gls{efc}}
In \gls{efc}, the confounder is provided as a non-bijective function $h$ of the
pre-outcome variables $\mb{t}$.
To reflect this property, we use $h(\mb{t})$ to denote the confounder.
As an illustrative example, let $\mathcal{G}$ be the Gamma distribution and consider $\mb{z}\in\{-1,1\}, p(\mb{z}=1) = 0.5$ is the confounder and the intervention of interest is $\mb{t} = \mb{z}*\mathcal{G}(1,\exp(\mb{z}))$.
Note $\textrm{sign}(\mb{t}) = \mb{z}$ meaning that $h(\mb{t}) = \textrm{sign}(\mb{t})$ is the confounder.
\Cref{fig:clarify-ht} shows causal graphs connecting our \gls{efc} notation to that in traditional \gls{ci}\xspace.
\begin{figure}[t]
\centering
\begin{subfigure}[b]{0.3\textwidth}
\begin{tikzpicture}
\node[state] (2) {$\mb{t}$};
\node[state] (3) [right =of 2] {$\mb{y}$};
\node[state] (1) [above right =of 2,xshift=-0.7cm,yshift=-0.7cm] {$\mb{z}$};
\path (1) edge node[above] {} (2);
\path (2) edge node[above] {} (3);
\path (1) edge node[above] {} (3);
\end{tikzpicture}
\caption{Traditional \gls{ci}\xspace} \label{fig:M1}
\end{subfigure}
\begin{subfigure}[b]{0.3\textwidth}
\begin{tikzpicture}
\node[state] (2) {$\mb{t}$};
\node[state] (3) [right =of 2] {$\mb{y}$};
\node[state] (1) [above right =of 2,xshift=-0.7cm,yshift=-0.7cm] {$h(\mb{t})$};
\path (1) edge node[above] {} (2);
\path (2) edge node[above] {} (3);
\path (1) edge node[above] {} (3);
\end{tikzpicture}
\caption{\gls{efc}} \label{fig:M2}
\end{subfigure}
\begin{subfigure}[b]{0.3\textwidth}
\begin{tikzpicture}
\node[state] (2) {$\mb{t}$};
\node[state] (3) [right =of 2] {$\mb{y}$};
\node[state] (1) [above right =of 2,xshift=-0.7cm,yshift=-0.7cm] {$h(\mb{t})$};
\path (2) edge node[above] {} (3);
\path (1) edge node[above] {} (3);
\end{tikzpicture}
\caption{Intervening in \gls{efc}} \label{fig:M2}
\end{subfigure}
\caption{Causal Graphs for Traditional \gls{ci}\xspace\ vs. \gls{efc}.}
\label{fig:clarify-ht}
\vspace{-0.4cm}
\end{figure}
With noise $\mb{\eta} \sim p(\mb{\eta}) (\mb{\eta}\rotatebox[origin=c]{90}{$\models$} \mb{t})$, our causal model samples, in order, the confounder "part" of pre-outcome variables $h(\mb{t})$, the pre-outcome variables $\mb{t}$, and the outcome $\mb{y}$ via the \emph{outcome function} $f$
\footnote{We also assume no interference~\cite{hernan2020causal} (also called Stable Unit Treatment Value Assumption~\cite{rubin1980randomization}) which means that an individual's outcome does not depend on others' treatment.
In \gls{efc}, when $\mb{t}$ and $\mb{\eta}$ are sampled IID there is no interference.
To see this, note $\forall i,j\,\, (\mb{t}_i, \mb{\eta}_i)\rotatebox[origin=c]{90}{$\models$} (\mb{t}_j, \mb{\eta}_j) \implies (\mb{y}_i ,\mb{t}_i) \rotatebox[origin=c]{90}{$\models$} (\mb{y}_j, \mb{t}_j) \implies \mb{y}_i \rotatebox[origin=c]{90}{$\models$} \mb{t}_j$.}:
\[h(\mb{t}) \sim p(h(\mb{t})) \quad \mb{t}~\sim p(\mb{t} \,\vert\, h(\mb{t})) \quad \mb{y} = f(\mb{t}, h(\mb{t}), \mb{\eta})\]
Similar to traditional \gls{ci}\xspace, for an intervention $\mathtt{t}^*$ the average effect, $\tau(\cdot)$, and the conditional effect, $\phi(\cdot, \cdot)$ at $h(\mathtt{t}^*_2)$, respectively, are defined as:
\begin{align}\label{eq:full-effs}
\quad \tau(\mathtt{t}^*)=\mathop{\E}_{h(\mb{t}), \mb{\eta}}[f(\mathtt{t}^*, h(\mb{t}),\mb{\eta})],
\quad \quad \quad \phi(\mathtt{t}^*,h(\mathtt{t}^*_2))=\mathop{\E}_{\mb{\eta}}[f(\mathtt{t}^*, h(\mathtt{t}^*_2),\mb{\eta})].
\end{align}
As the pre-outcome variables determine the confounder, positivity is violated.
Further, the \emph{outcome function} $f(\mb{t}, h(\mb{t}), \mb{\eta})$ could recover the exact value of $h(\mb{t})$ from $\mb{t}$ instead of its second argument.
Thus, two different outcome functions could lead to the same observational data distribution, posing a fundamental obstacle to causal effect estimation.
This is the central challenge in \gls{efc}.
\subsection{Causal Questions With Functional Positivity}\label{sec:qs-without-pos}
Without positivity, we can only estimate the effects of certain functions of $\mb{t}$.
We call such interventions, on some function $g(\mb{t})$, \emph{functional interventions}.
The implied causal model for the outcome for functional intervention
value $g(\mathtt{t}^*)$ and confounder value $h(\mathtt{t}^*_2)$
is first $\mb{t} \sim p(\mb{t} \,\vert\, g(\mb{t}) = g(\mathtt{t}^*), h(\mb{t}) = h(\mathtt{t}^*_2))$ and then $\mb{y} = f(\mb{t}, h(\mathtt{t}^*_2), \mb{\eta})$
\footnote{Intervening on $g(\mb{t})$ can be interpreted as making a \emph{soft intervention}~\citep{eberhardt2007interventions,correa2020calculus} of $\mb{t}$ to $p(\mb{t}\,\vert\, \mb{z}, g(\mb{t})=g({\tilde{\mathtt{t}}}))$.
}.
Then, the \textit{functional} average effect is
\begin{align*}
(\text{average})\quad &\tau(g(\mathtt{t}^*))={\E}_{h(\mb{t}), \mb{\eta}}\E_{\mb{t} \,\vert\, g(\mb{t}) = g(\mathtt{t}^*), h(\mb{t})}[f(\mb{t}, h(\mb{t}), \mb{\eta})].
\end{align*}
An example of a functional intervention is intervening on the cumulative dosage of a drug.
In contrast, traditional interventions would set each individual dose given at different points in time.
\paragraph{\gls{func-pos} and Functional Effect Estimation}
For the causal model above to be well-defined for all functional interventions~$g(\mathtt{t}^*)$, the conditional $p(\mb{t} \,\vert\, g(\mb{t}) = g(\mathtt{t}^*), h(\mb{t}) = h(\mathtt{t}^*_2))$ must exist.
To guarantee this existence, we define \emph{\glsreset{func-pos}\gls{func-pos}} for any $g(\mathtt{t}^*)$
\begin{align}\label{eq:functional-positivity}
(\text{\gls{func-pos}})\quad p(h(\mb{t}) = h(\mathtt{t}^*_2))>0 \Longrightarrow p(g(\mb{t}) = g(\mathtt{t}^*) \,\vert\, h(\mb{t}) = h(\mathtt{t}^*_2)) > 0 .
\end{align}
\Gls{func-pos} says that the function of the pre-outcome variables that is being intervened on needs to have sufficient randomness when the function of the pre-outcome variables that defines the confounders is fixed.
Further, under \gls{func-pos}, effect estimation for functional interventions\ is reduced to traditional \gls{ci}\xspace\ on data $p(\mb{y}, g(\mb{t}), h(\mb{t}))$.
With positivity and ignorability satisfied, traditional causal estimators such as propensity scores \citep{rosenbaum1983central}, matching \citep{ratkovic2014balancing}, regression~\citep{hill2010bayesianci}, and doubly robust methods \citep{robins2000robust} can be used to estimate the causal effect.
Focusing on regression, let $f_\theta$ be a flexible function, then
$\min_{\theta} \E_{y, \mb{t}} [(\mb{y} - f_\theta(h(\mb{t}), g(\mb{t})))^2]$
would estimate the conditional expectation of interest : $\E[\mb{y}\,\vert\, h(\mb{t}), g(\mathtt{t}^*)]$.
With $\theta$, the effect of $g(\mathtt{t}^*)$ can be estimated by averaging the estimate of the conditional expectation over the marginal distribution $p(h(\mb{t}))$:
\begin{align}
\tau(g(\mathtt{t}^*)) = \E_{\mb{t}}[f_\theta(h(\mb{t}), g(\mathtt{t}^*))].
\label{eq:causal-estimate-positivity}
\end{align}
\section{Identification of effects of the full intervention}\label{sec:full-est}
When positivity is violated, causal effects cannot be estimated as conditional expectations over the observed data in general.
We give a functional condition, called \glsreset{c-red}\gls{c-red}, that allows us to estimate the effect of the full intervention $do(\mb{t}=\mathtt{t}^*)$, even when positivity is violated.
Specifically, \gls{c-red} allows us to construct a \textit{surrogate intervention} $\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))$ whose conditional effect at $h(\mathtt{t}^\prime)$ matches the conditional effect of interest, $\phi(\mathtt{t}^*, h(\mathtt{t}^*_2))$.
Let ${\tilde{\mathtt{t}}}$ be a fixed value of the full intervention, then \gls{c-red} is
\begin{assumption*}\label{assn:c-redundancy}
Recall the outcome $y = f({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}}), \eta)$. With $\nabla_{\tilde{\mathtt{t}}}$ as gradient w.r.t.\ to argument ${\tilde{\mathtt{t}}}$:
\[\forall {\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}}_2), \eta, \quad \nabla_{{\tilde{\mathtt{t}}}} f ({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}}_2), \eta)^T\nabla_{{\tilde{\mathtt{t}}}} h({\tilde{\mathtt{t}}}) = 0.\]
\end{assumption*}
In words, \gls{c-red} is the condition that the outcome function $f$ uses the value of the confounder from its second argument instead of computing $h(\mb{t})$ from the first argument\footnote{If $f$ transforms its first argument ${\tilde{\mathtt{t}}}$ into $h({\tilde{\mathtt{t}}})$ as one amongst many different computations, the chain rule implies $\nabla_{\tilde{\mathtt{t}}} f({\tilde{\mathtt{t}}}, h(\mathtt{t}^*_2))^\top \nabla_{\tilde{\mathtt{t}}} h({\tilde{\mathtt{t}}})$ has a term $\|\nabla_{\tilde{\mathtt{t}}} h({\tilde{\mathtt{t}}})\|^2 $ which is non-zero in general.}.
To compute the conditional effect $\phi(\mathtt{t}^*, h(\mathtt{t}^*_2))$, we develop \glsreset{ours}\gls{ours}.
\gls{ours}'s key step is to construct a surrogate intervention $\mathtt{t}^\prime(\mathtt{t}^*,h(\mathtt{t}^*_2))$ such that
\begin{align*}
\begin{split}
\phi(\mathtt{t}^*, h(\mathtt{t}^*_2)) &= \phi({\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))}, h(\mathtt{t}^*_2)), \quad \quad
h(\mathtt{t}^*_2) = h(\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))) .
\end{split}
\end{align*}
\begin{wrapfigure}[9]{r}{0.32\textwidth}
\vspace{-25pt}
\centering
\includegraphics[width=0.3\columnwidth]{images/level-sets-final}
\caption{\gls{ours}'s traversal.}
\label{fig:level-sets}
\end{wrapfigure}
By definition, a surrogate intervention lives in the conditional effect level-set: $\{{\tilde{\mathtt{t}}}: \phi({\tilde{\mathtt{t}}}, h(\mathtt{t}^*_2))= \phi(\mathtt{t}^*, h(\mathtt{t}^*_2))\}$.
So \gls{ours} searches this level-set for $\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))$.
See~\cref{fig:level-sets} which plots the conditional effect level-sets with the value of $h(\mb{t})$ fixed (\textcolor{red}{red}) in $(\textrm{supp}(\mb{t}), \textrm{supp}(h(\mb{t})))$-space.
\textcolor{green}{Green} corresponds to the observed data, $\textrm{supp}(\mb{t}, h(\mb{t}))$.
\gls{ours} finds $\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))$ by traversing the level-sets (\textbf{black}) to account for the confounder part mismatch $h(\mathtt{t}^*)\not=h(\mathtt{t}^*_2)$.
\gls{c-red} ensures \gls{ours} can traverse these level-sets as it implies $\nabla_{{\tilde{\mathtt{t}}}}\phi({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}}_2))\nabla_{{\tilde{\mathtt{t}}}}h({\tilde{\mathtt{t}}}) = 0$ under the regularity conditions in~\cref{thm:thm-c-red}.
Thus, under \gls{c-red}, surrogate interventions can be constructed by solving a gradient flow equation which guarantees identification as follows:
\newcommand{\lemmaone}{
Assume \gls{c-red} holds. Assuming the following:
\begin{enumerate}[leftmargin=0.5cm]
\item Let $\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))$ be the limiting solution to the gradient flow equation $\fullder{{\tilde{\mathtt{t}}}(s)}{s} = -\nabla_{{\tilde{\mathtt{t}}}}(h({\tilde{\mathtt{t}}}(s)) - h(\mathtt{t}^*_2))^2$, initialized at ${\tilde{\mathtt{t}}}(0)=\mathtt{t}^*$; i.e. $\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2)) = \lim_{s\rightarrow \infty } {\tilde{\mathtt{t}}}(s)$.\\
Further, let $h(\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))) = h(\mathtt{t}^*_2)$
and $\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))\in \textrm{supp}(\mb{t})$.
\item $f({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}}), \eta)$ and $h({\tilde{\mathtt{t}}})$ as functions of ${\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}})$ are continuous and differentiable and the derivatives exist for all ${\tilde{\mathtt{t}}},\eta$.
Let $\nabla_{{\tilde{\mathtt{t}}}} f({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}}), \eta)$ exist and be bounded and integrable w.r.t.\ the probability measure corresponding to $p(\mb{\eta})$, for all values of ${\tilde{\mathtt{t}}}$ and $h({\tilde{\mathtt{t}}})$.
\end{enumerate}
Then the conditional effect (and therefore the average effect) is identified:
\begin{align}\label{eq:effect-estimates}
\begin{split}
\phi(\mathtt{t}^*,h(\mathtt{t}^*_2)) = \phi\left(\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2)), h(\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2)))\right)= \E\left[\mb{y}\,\vert\, \mb{t} = {\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))}\right]
\end{split}
\end{align}
}
\begin{theorem}\label{thm:thm-c-red}
\lemmaone{}
\end{theorem}
In words, the key idea is that starting at ${\tilde{\mathtt{t}}}(0) = \mathtt{t}^*$ and following $\nabla_{\tilde{\mathtt{t}}} h({\tilde{\mathtt{t}}})$ means ${\tilde{\mathtt{t}}}(s)$ always lies in the level-set $\{{\tilde{\mathtt{t}}} : \phi({\tilde{\mathtt{t}}}, h(\mathtt{t}^*_2)) = \phi(\mathtt{t}^*, h(\mathtt{t}^*_2))\}$.
See~\cref{appsec:thm-c-red} for the proof.
While \gls{c-red} is stated in terms of the gradient of the outcome function, it suffices for~\cref{thm:thm-c-red} to assume a weaker condition about the gradient of the conditional effect: $\nabla_{\tilde{\mathtt{t}}} \E_{\mb{\eta}} f({\tilde{\mathtt{t}}}, {\tilde{\mathtt{t}}}_2, \mb{\eta})^\top \nabla_{{\tilde{\mathtt{t}}}} h({\tilde{\mathtt{t}}}) = 0$.
\paragraph{Surrogate Positivity}
In~\cref{thm:thm-c-red}, we assumed that the surrogate $\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2)) \in \textrm{supp}(\mb{t})$.
This condition, which we call surrogate positivity (analogous to positivity), states that for any intervention and confounder, surrogate interventions that are limiting solutions to the gradient flow equation have nonzero density conditional on the confounder value.
Formally, for any intervention $\mb{t} =\mathtt{t}^*$
\begin{align}\label{eq:surr-pos}
p(h(\mb{t}) = h(\mathtt{t}^*_2)) > 0 \implies p(\mb{t} = \mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2)) \,\vert\, h(\mb{t}) = h(\mathtt{t}^*_2)) > 0,
\end{align}
and $\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))$ satisfies assumption 1 in~\cref{thm:thm-c-red}.
Surrogate positivity along with \gls{c-red}, is sufficient for full effect estimation under \gls{efc}.
Next, we show that the positivity assumption in traditional causal inference is a special case of surrogate positivity.
\paragraph{\glsreset{ci}Traditional \gls{ci}\xspace and \gls{ours}}
Let the confounder and intervention of interest in traditional \gls{ci}\xspace be $\mb{z}$ and $\mb{a}$ respectively. Assume both are scalars and ignorability and positivity hold.
This setup can be embedded in \gls{efc} by defining the vector of pre-outcome variables as:
$
\mb{t} = [\mb{a}; \mb{z}].
$
In this setting, \gls{c-red} and surrogate positivity(\cref{eq:surr-pos}) hold by default.
Let the outcome be $\mb{y} = f(\mb{t}, h(\mb{t})) = f(\mb{a}, \mb{z})$, where $f$ only depends on the first element of $\mb{t}$, i.e.\ $\mb{a}$\footnote{We ignore noise in the outcome for ease of exposition.}.
Let $e_1 = [1, 0]$ and $e_2 = [0, 1]$.
In traditional \gls{ci}\xspace as \gls{efc}, $\nabla_{{\tilde{\mathtt{t}}}} f({\tilde{\mathtt{t}}}, h(\mathtt{t}^*_2)) \propto e_1$ and $\nabla_{{\tilde{\mathtt{t}}}} h({\tilde{\mathtt{t}}}) \propto e_2$ meaning that $\nabla_{{\tilde{\mathtt{t}}}} f({\tilde{\mathtt{t}}}, h(\mathtt{t}^*_2)) ^\top \nabla_{{\tilde{\mathtt{t}}}} h({\tilde{\mathtt{t}}}) = 0$.
Thus, \gls{c-red} holds by default.
Moreover, under positivity of $\mb{a}$ w.r.t.\ $\mb{z}$, we also have surrogate positivity for traditional \gls{ci}\xspace as an \gls{efc} problem.
In this setting, \gls{ours} computes $\mathtt{t}^\prime=[a^*, h(\mathtt{t}^*_2)]$ by following $-\nabla_{{\tilde{\mathtt{t}}}} h({\tilde{\mathtt{t}}}) = [0, -1]$, which only changes the value of $h({\tilde{\mathtt{t}}}_2)$, not the value of $a$.
Thus, $\mathtt{t}^*$ and $\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))$ will have the same first element and $\mathtt{t}^\prime$'s second element will be $h(\mathtt{t}^*_2)$. As $\mb{a}$ has positivity w.r.t.\ $\mb{z}$, we have $p(\mb{a} = a^* , \mb{z}=h(\mathtt{t}^*_2))>0$ which means $\mathtt{t}^\prime\in\textrm{supp}(\mb{t})$. The estimated conditional effect is $\E[\mb{y} \,\vert\, \mb{t} = \mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))] = f([a^*, z^*], h(\mathtt{t}^*_2)) = \E[\mb{y} \,\vert\, \mb{a} = a^*, \mb{z}=h(\mathtt{t}^*_2))]$, which matches the estimate in traditional \gls{ci}\xspace.
\paragraph{Implementation of \gls{ours}}
\Gls{ours} first estimates the conditional expectation $\E[\mb{y}\,\vert\, \mb{t}]$; this can be done with model-based or nonparametric estimators.
This is achieved by regressing $\mb{y}$ on $\mb{t}$, $\hat{f} = \argmin_{u\in \mathcal{F}} \E_{\mb{y}, \mb{t} \sim D} (\mb{y} - u(\mb{t}))^2 $, with empirical distribution $D$.
The surrogate intervention $\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))$ is computed using Euler integration to solve the gradient flow equation.
Euler integration in this setting is equivalent to gradient descent with a fixed step size.
Other, more efficient schemes like Runge–Kutta numerical integration methods~\cite{ascher1998computer} could also be used.
The conditional effect estimate is $\hat{f}(\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2)))$. See~\cref{alg:full-eff-estimation} for a description.
\subsection{Estimation error of \gls{ours} in practice}\label{sec:estimation-error}
To compute the surrogate intervention $\mathtt{t}^\prime$, \gls{ours} uses the gradients of $h(\cdot)$ in Euler integration.
In practice, taking Euler integration steps, instead of solving the gradient flow exactly, could result in errors.
Then $\mathtt{t}^\prime$ could lie outside the level-set of the conditional effect $\phi(\mathtt{t}^*, h(\mathtt{t}^*_2)) = \E_{\mb{\eta}}[f(\mathtt{t}^*, h(\mathtt{t}^*_2), \mb{\eta})]$.
Further, if $h(\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2)))\not = h(\mathtt{t}^*_2)$, \gls{ours} incurs error for conditioning on a value of the confounder that is different from $h(\mathtt{t}^*_2)$.
The error due to $\mathtt{t}^\prime$ estimation is decoupled from the error in the estimation of $\E[\mb{y} \,\vert\, \mb{t}]$ which adds without further amplification.
We formalize this error:
\newcommand{\theoremmain}{
Consider the conditional effect $ \phi(\mathtt{t}^*, h(\mathtt{t}^*_2))$.
Let $\hat{\mathtt{t}}(\mathtt{t}^*, h(\mathtt{t}^*_2))$ be the estimate of the surrogate intervention computed by \gls{ours}, computed via Euler integration of the gradient flow $\fullder{{\tilde{\mathtt{t}}}(s)}{s} = - \nabla_{\tilde{\mathtt{t}}} (h({\tilde{\mathtt{t}}}(s)) - h(\mathtt{t}^*_2))^2$, initialized at ${\tilde{\mathtt{t}}}(0) = \mathtt{t}^*$.
Assume the true surrogate $\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))$ exists and is the limiting solution to the gradient flow equation.
\begin{enumerate}
\item Let the finite sample estimator of $\E[\mb{y}\,\vert\, \mb{t}={\tilde{\mathtt{t}}}]$ be $\hat{f}({\tilde{\mathtt{t}}})$.
Let
the error for all ${\tilde{\mathtt{t}}}$ be bounded, $|\hat{f}({\tilde{\mathtt{t}}}) - \E[\mb{y}\,\vert\, \mb{t}={\tilde{\mathtt{t}}}] |\leq c(N)$, where $N$ is the sample size and $ \lim_{ N\rightarrow \infty } c(N) = 0 $.
\item Assume $K$ Euler integrator steps were taken to find the surrogate estimate $\hat{\mathtt{t}}(\mathtt{t}^*, h(\mathtt{t}^*_2))$, each of size $\ell$.
Let the maximum confounder mismatch be $\max_{i\leq K}(h({\tilde{\mathtt{t}}}_i) - h(\mathtt{t}^*_2))^2 = M$.
\item Let $L_{z,{\tilde{\mathtt{t}}}}$ be the Lipschitz-constant of $\phi({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}}_2))$ as a function of $h({\tilde{\mathtt{t}}}_2)$, for fixed ${\tilde{\mathtt{t}}}$.\\
Let $L_e$ be the Lipschitz-constant of $\E[\mb{y} \,\vert\, \mb{t}={\tilde{\mathtt{t}}}] = \phi({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}}))$ as a function of ${\tilde{\mathtt{t}}}$.
\\
Assume $h$ has a gradient with bounded norm, $\|\nabla h({\tilde{\mathtt{t}}})\|_2 < L_h$.
\\
Assume $f$'s Hessian has bounded eigenvalues: $\forall {\tilde{\mathtt{t}}}, {\tilde{\mathtt{t}}}_2,\,\, \| \nabla^2_{\tilde{\mathtt{t}}} \phi({\tilde{\mathtt{t}}}, h({\tilde{\mathtt{t}}}_2))\|_2 \leq \sigma_{\mathtt{H} \phi} $.
\end{enumerate}
The conditional effect estimate error, $\xi(\mathtt{t}^*,h(\mathtt{t}^*_2)) = |\hat{f}(\hat{\mathtt{t}}) - \phi(\mathtt{t}^*, h(\mathtt{t}^*_2))|$, is upper bounded by:
\begin{align}\label{eq:est-err-bound}
c(N) + \min\left(L_e\|\mathtt{t}^\prime - \hat{\mathtt{t}}\|_2,
\,\, 2K \ell^2 \left(\mathcal{O}(\ell) + M \sigma_{\mathtt{H} \phi } L^2_{h} \right) + L_{z,\hat{\mathtt{t}}} \|h(\hat{\mathtt{t}}) - h(\mathtt{t}^*_2)\|_2\right)
\end{align}
}
\begin{theorem}\label{thm:main-effect-bound}
\theoremmain
\vspace{2pt}
\end{theorem}
See \cref{appsec:lode-error} for the proof.
\Cref{thm:main-effect-bound} captures the trade-off between biases due to conditioning on the wrong confounder value and due to the accumulated error in solving the gradient flow equation.
This accumulated error analysis may be loose in settings where the sum of many gradient steps lead to $\hat{\mathtt{t}} \approx \mathtt{t}^\prime$, even if each step individually induces large error.
In such settings, the term that depends on $\|\hat{\mathtt{t}} - \mathtt{t}^\prime\|_2$ is a better measure of error.
The maximum-mismatch $M$ appears because Euler integrator takes steps that depend on the magnitude of the gradient which depends on the mismatch value $(h({\tilde{\mathtt{t}}}_i) - h(\mathtt{t}^*_2))$.
If mismatch is large for some $i$, the Euler step could lead to a large error for a fixed step size $\ell$.
We discuss the assumptions in~\cref{thm:thm-c-red,thm:main-effect-bound} in~\cref{appsec:assumptions}
\subsection{Effect Connectivity and the Existence of $\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))$}
The key element in \Cref{thm:thm-c-red} is the surrogate intervention $\mathtt{t}^\prime$ such that its conditional effect given $h(\mathtt{t}^\prime)$, equals that of $\mathtt{t}^*$ and $ h(\mathtt{t}^*_2)$.
The orthogonality $\nabla_{{\tilde{\mathtt{t}}}} f^\top \nabla_{{\tilde{\mathtt{t}}}} h =0 $, is a functional condition that does not guarantee $\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))$ exists in $\textrm{supp}(\mb{t})$; a necessity to compute $\E[\mb{y}\,\vert\, \mb{t} = \mathtt{t}^\prime]$ without additional parametric assumptions.
We give a general condition called \emph{Effect Connectivity} that guarantees the surrogate intervention exists.
With conditional effect $\phi(\mathtt{t}^*, h(\mathtt{t}^*_2))$, for any $\mathtt{t}^*$
\begin{align}\label{eq:weird-func-pos-condition}
p(h(\mb{t})=h(\mathtt{t}^*_2))>0 \implies p(\phi(\mb{t}, h(\mb{t})) = \phi(\mathtt{t}^*,h(\mathtt{t}^*_2))\,\vert\, h(\mb{t}) = h(\mathtt{t}^*_2)) > 0.
\end{align}
In words, $\mb{t}$ has a chance of setting the conditional effect to any possible value $\textrm{supp}( \phi(\mb{t}, h(\mb{t}_2)))$ given any confounder value $h(\mathtt{t}^*_2) \in \textrm{supp}(h(\mb{t}))$.
An equivalent statement is that every level set of the conditional effect $\phi(\mathtt{t}^*, h(\mathtt{t}^*_2))$, with $h(\mathtt{t}^*_2)$ fixed, contains an intervention for each confounder value.
That is, for some $h(\mathtt{t}^*_2)$ define the level set $A_c = \{\mathtt{t}^*; f(\mathtt{t}^*, h(\mathtt{t}^*_2))=c\}$, then $\forall h(\mathtt{t}^*_2)\in \textrm{supp}(h(\mb{t})),\, p(\mb{t} \in A_c \,\vert\, h(\mb{t}) =h(\mathtt{t}^*_2) ) > 0$.
\newcommand{\sufficiencylemma}{
Under Effect Connectivity, \cref{eq:weird-func-pos-condition}, any surrogate intervention $\mathtt{t}^\prime(\mathtt{t}^*,h(\mathtt{t}^*_2)) \in \textrm{supp}(\mb{t})$.
}
\begin{theorem}\label{thm:thm-sufficiency}
\sufficiencylemma{}
\end{theorem}
We give the proof in~\cref{appsubsec:sufficiency}.
Whether the intervention $\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))$ can be found via tractable search is problem-specific.
If the surrogate $\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))$ exists $\forall \mathtt{t}^*,h(\mathtt{t}^*_2)$, then \cref{eq:weird-func-pos-condition} holds by definition of the surrogate.
Effect Connectivity allows us to reason about values of $f$ anywhere in $\textrm{supp}(\mb{t})\times\textrm{supp}(h(\mb{t}))$ using only samples from $p(\mb{y}, \mb{t})$.
Further, it is necessary in \gls{efc}:
\newcommand{\necessitylemma}{
Effect Connectivity is necessary for nonparametric effect estimation in \gls{efc}.
}
\begin{theorem}\label{thm:thm-necessity}
\necessitylemma{}
\end{theorem}
We prove this in~\cref{appsubsec:necessity-proof}.
Effect Connectivity ensures that causal models with different causal effects have different observational distributions.
Then, parametric assumptions on the causal model are not necessary to estimate effects.
\input{sections/experiments.tex}
\section{Discussion}
When positivity is violated in traditional \gls{ci}\xspace, not all effects are estimable without further assumptions.
In such cases, practitioners have to turn to parametric models to estimate causal effects.
However, parametric models can be misspecified when used without underlying causal mechanistic knowledge.
We develop a new general setting of observational causal effect estimation called \glsreset{efc}\gls{efc} where the confounder can be expressed as a function of the data, meaning positivity is violated.
Even when positivity is violated, the effects of many functional interventions{} are estimable.
We develop a sufficient condition called \glsreset{func-pos}\gls{func-pos}\ to estimate effects of functional interventions.
Such effects could be of independent interest; like the effect of cumulative dosage of a drug instead of joint effects of multiple dosages at different times.
Second, we prove a necessary condition for nonparametric estimation of effects of the full intervention.
We propose the \gls{c-red} condition, under which, the effect of the full intervention on $\mb{t}$ is estimable without parametric restrictions.
We develop \glsreset{ours}\gls{ours} that computes surrogate interventions whose effects are estimable and match a conditional effect of interest.
Further, we give bounds on errors (\cref{thm:main-effect-bound}) induced due to imperfect estimation of the surrogate intervention.
Finally, we empirically demonstrate \gls{ours}'s ability to correct for confounding in both simulated and real data.
\paragraph{Future.} A few directions of improvement remain which we elaborate next.
First, \gls{func-pos} may not hold for all functions $g(\mb{t})$ that we want to intervene on.
Instead, one could compute a ``projection'' $g_\Pi$ to the space of functions that satisfy \gls{func-pos} and inspect the effects defined by $g_\Pi$ instead.
A second direction of interest is to let $h(\mb{t})$ only account for a part of the confounding, meaning ignorability is violated.
This bias could be mitigated under smoothness conditions of the outcome function and its interaction with the degree of violation of ignorability.
Finally, \gls{ours}'s search strategy is Euler integration, which is equivalent to gradient descent with a fixed step size.
Optimization techniques like momentum, rescaling the gradient using an adaptive matrix, and using second order hessian information, speed up gradient descent.
However, if there are many local or global minima for $(h({\tilde{\mathtt{t}}}) - h(\mathtt{t}^*_2))^2$, such techniques will result in a different solution than Euler integration, which could mean that effect estimates are biased.
One extension of \gls{ours} would allow for search strategies that use such techniques.
\newpage
\section*{Broader Impact}
Our work mainly applies to causal inference where confounders are specified as functions of observed data, such as in problems in genetics and healthcare.
We choose to assess the impact of our work through its applications in these fields.
A positive impact of the work is that better estimates of causal effects helps guide treatment for people and aid in understanding biological pathways of diseases.
However, in healthcare, data collected in hospitals has biases.
If, for instance, a certain demographic of people have more complete data collected about them, then this demographic would have better quality effect estimates, potentially meaning that they receive better treatment.
This problem could be characterized by evaluating the positivity of treatment and completeness of confounders in electronic health record data split by demographics.
\section*{Acknowledgements}
The authors were partly supported by NIH/NHLBI Award R01HL148248, and by NSF Award 1922658 NRT-HDR: FUTURE Foundations, Translation, and Responsibility for Data Science.
The authors would like to thank Xintian Han, Raghav Singhal, Victor Veitch, Fredrik D. Johansson and the reviewers for thoughtful feedback.
The authors would also like to thank Mukund Sudarshan and Prof. Sriram Sankararaman for help with running the \gls{gwas} experiments.
\section{Experiments}\label{sec:exps}
We evaluate \gls{ours} on simulated data first and show that \gls{ours} can correct for confounding.
We also investigate the error induced by imperfect estimation of the surrogate intervention in \gls{ours}.
Further, we run \gls{ours} on a \gls{gwas} dataset~\citep{wellcome2007genome} and demonstrate that \gls{ours} is able to correct for confounding and recovers genetic variations that have been reported relevant to Celiac disease~\citep{dubois2010multiple,sollid2002coeliac,hunt2008novel,adamovic2008association}.
\subsection{Simulated experiments}
We investigate different properties of \gls{ours} on simulated data where ground truth is available.
Let the dimension of $\mb{t}$ (pre-outcome variables) be $T=20$ and outcome noise be $\mb{\eta} \sim \mathcal{N}(0,0.1)$.
We consider two \gls{efc} causal models, denoted by $A$ and $B$ with different $h(\mb{t})$ and $f(\mb{t}, h(\mb{t}), \mb{\eta})$:
\begin{align*}
(A) \quad & h(\mb{t}) = \gamma \frac{\sum_i \mb{t}_i}{\sqrt{T}},\quad \quad \mb{t}\sim \mathcal{N}(0, \sigma^2\mathbb{I}^{T\times T}),
\quad y = \frac{\sum_i(-1)^i \mb{t}_i}{\sqrt{T}} + \alpha h(\mb{t})^2 + (1+\alpha)h(\mb{t}) + \mb{\eta}\\
(B) \quad & h(\mb{t}) = \mathop{\textstyle \sum}_{ i:i \in 2\mathbb{Z}} \gamma \mb{t}_i\mb{t}_{i+1},\quad \mb{t}\sim \mathcal{N}(0, \sigma^2\mathbb{I}^{T\times T}),
\quad y = \frac{\sum_i (-1)^i \mb{t}^2_i}{\sqrt{T}} + \alpha h(\mb{t})+ \mb{\eta}
\end{align*}
In both causal models, \gls{c-red} is satisfied.
The constant $\gamma$ controls the strength of the confounder and the constant $\alpha$ controls the Lipschitz constant of the outcome as a function of the confounder.
We let the variance $\sigma^2 = 1$, unless specified otherwise.
In the following, we train on $1000$ samples and report conditional effect \gls{rmse}, computed with another $1000$ samples.
We used a degree-2 kernel ridge regression to fit the outcome model as a function of $\mb{t}$.
This model is correctly specified, and so the conditional $\E[\mb{y}\,\vert\, \mb{t} = {\tilde{\mathtt{t}}}]$ can be estimated well.
We compare against a baseline estimate of conditional effect that is the same outcome model's estimate of $\E[\mb{y} \,\vert\, \mb{t}=\mathtt{t}^*]$.
This baseline fails to account for confounding and produces a biased estimate of the conditional effect of $do(\mb{t}=\mathtt{t}^*)$, conditional on any $h(\mathtt{t}^*_2)\not=h(\mathtt{t}^*)$.
\begin{figure*}[t!]
\centering
\begin{subfigure}[b]{0.45\textwidth}
\centering
\includegraphics[width=\textwidth]{images/lode-efc-numt20-quad}
\caption{Causal Model $A$}
\end{subfigure}
~
\begin{subfigure}[b]{0.45\textwidth}
\centering
\includegraphics[width=\textwidth]{images/lode-efc-numt20-nonlin}
\caption{Causal Model $B$}
\end{subfigure}
\caption{\Gls{rmse} of estimated \condeff{} vs. strength of confounding $\gamma$.
\gls{ours} corrects for confounding and produces good effect estimates across different values of $\gamma$.}
\label{fig:lode-est}
\end{figure*}
First, we investigate how well \gls{ours} can correct for confounding for both causal models.
We let $\alpha=1$ and obtain surrogate estimates by Euler integrating until the quantity $\E_{\mathtt{t}^*, h(\mathtt{t}^*_2)}(h({\tilde{\mathtt{t}}}(s))- h(\mathtt{t}^*_2))^2$ is smaller than $10^{-4}$ times value at initialization, where $\E_{\mathtt{t}^*, h(\mathtt{t}^*_2)}$ is expectation over the evaluation set.
In~\cref{fig:lode-est}, we plot the mean and standard deviation of conditional effect \gls{rmse} averaged over $10$ seeds, for different strengths of confounding.
We see that \gls{ours} is able to estimate effects well across multiple strengths of confounding while the baseline suffers.
\begin{figure}[t]
\centering
\begin{subfigure}[b]{0.45\textwidth}
\centering
\includegraphics[width=\textwidth]{images/variance-error-lin-new}
\caption{Causal Model $A$}
\end{subfigure}
~
\begin{subfigure}[b]{0.45\textwidth}
\centering
\includegraphics[width=\textwidth]{images/variance-error-nonlin-new}
\caption{Causal Model $B$}
\end{subfigure}
\caption{\gls{rmse} of estimated \condeff{} estimate vs. the strength of confounding $\gamma$, for different levels of variance of $\mb{t}$, $\sigma^2$.
Small $\sigma$ leads to large conditional estimation error.}
\label{fig:eff-pos-experiment}
\end{figure}
\begin{wrapfigure}[14]{r}{0.42\textwidth}
\centering
\includegraphics[width=0.42\textwidth]{images/lode-non-perf-v-LR}
\caption{
\Gls{rmse} of estimated \condeff{} vs. step size in Euler Integrator in causal model $B$.
Accumulating error due to large step size in Euler integrator increases with strength of confounding.
}
\label{fig:lr-error}
\end{wrapfigure}
Second, we investigate \gls{ours}'s estimation when surrogate positivity holds but the probability $p(\mb{t} \approx \mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2)))$ is very small.
This results in estimation error due to poor fitting of the outcome model in low density regions of $\textrm{supp}(\mb{t})$.
We run \gls{ours} on simulated data where $\mb{t}$ is generated with different variances ($\sigma^2$).
For small $\sigma$, the outcome model error is large when using surrogate interventions $\mathtt{t}^\prime(\mathtt{t}^*,h(\mathtt{t}^*_2))$, where either $h(\mathtt{t}^*_2)$ or $\mathtt{t}^*$ is large.
This leads to high variance effect estimation as we show in ~\cref{fig:eff-pos-experiment} for both causal models.
For various variances of $\mb{t}$, $\sigma^2$, we plot the mean and standard deviation of \gls{rmse} of estimated conditional effect over $10$ seeds, against different $\gamma$.
Third, we investigate the bias induced due to imperfect estimation of the surrogate intervention in \gls{ours} for both causal models.
We construct surrogate interventions $\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))$ by ensuring there is confounder-value mismatch $h({\tilde{\mathtt{t}}})\not = h(\mathtt{t}^*_2)$. We do this by interrupting Euler integration when the objective
$\mathop{\E}_{\mathtt{t}^*, h(\mathtt{t}^*_2)}(h(\mathtt{t}^\prime(\mathtt{t}^*, h(\mathtt{t}^*_2))) - h(\mathtt{t}^*_2))^2 = \delta^2 > 0,$
where the $\E_{\mathtt{t}^*, h(\mathtt{t}^*_2)}$ is over our evaluation set upon which we estimate conditional effects.
For different $\alpha$, we plot in~\cref{fig:est-err-confmismatch} the mean and standard deviation of \gls{rmse} of estimated conditional effect over $10$ seeds, against different degrees of confounder mismatch, $\delta$.
The error due to confounder mismatch is mitigated by small $\alpha$, the Lipschitz-constant of the outcome as a function of $h(\mb{t})$.
\begin{figure*}[t]
\centering
\begin{subfigure}[b]{0.45\textwidth}
\centering
\includegraphics[width=\textwidth]{images/lode-efc-numt20-conferr}
\caption{Causal Model $A$}
\end{subfigure}
~
\begin{subfigure}[b]{0.45\textwidth}
\centering
\includegraphics[width=\textwidth]{images/lode-efc-numt20-conferr-nonlin}
\caption{Causal Model $B$}
\end{subfigure}
\caption{\Gls{rmse} of estimated \condeff{} vs. degree of confounder mismatch $\delta$.
Error due to conditioning on a mismatched value of the confounder increases with strength of confounding but is mitigated by smoothness of the outcome function. }
\label{fig:est-err-confmismatch}
\end{figure*}
Finally, we consider how step size in Euler integration affects the quality of estimated effects.
Large step sizes may result in biased surrogate estimates; this bias is captured in the accumulation error in~\cref{sec:estimation-error}.
We focus on the non-linear case in causal model B where gradient errors can accumulate(see~\cref{appsec:linear-conf-func}).
We demonstrate this error in~\cref{fig:lr-error} where we plot mean and standard deviation of conditional effect \gls{rmse} against the strength of confounding, for different step sizes $\ell$.
We do not report results for larger step sizes ($\ell>2$) because Euler integration diverged for many surrogate estimates.
\subsection{Effects in Genetics (\gls{gwas})}
In this experiment, we explore the associations of genetic factors and Celiac disease.
We utilize data from the Wellcome Trust Celiac disease \gls{gwas} dataset~\citep{dubois2010multiple,wellcome2007genome} consisting of individuals with celiac disease, called cases $(n = 3796)$, and controls $(n = 8154)$.
We construct our dataset by filtering from the $\sim 550,000$ \glspl{snp}.
The only preprocessing in our experiments is linkage disequilibrium pruning of adjacent \glspl{snp} (at $0.5$ $R^2$) and PLINK~\cite{chang2015second} quality control.
After this, $337,642$ \glspl{snp} remain for $11,950$ people.
We imputed missing \glspl{snp} for each person by sampling from the marginal distribution of that \gls{snp}.
No further \gls{snp} or person was dropped due to missingness.
The objective of this experiment is to show that \gls{ours} corrects for confounding and recovers \glspl{snp} reported in the literature~\citep{dubois2010multiple,sollid2002coeliac,hunt2008novel,adamovic2008association}.
To this end, after preprocessing, we included in our data $50$ \glspl{snp} reported in ~\cite{dubois2010multiple,sollid2002coeliac,hunt2008novel,adamovic2008association} and $1000$ randomly sampled from the rest.
We use outcome models and functional confounders $h()$ traditionally employed in the \gls{gwas} literature.
We choose a linear $h({\tilde{\mathtt{t}}}) = A^\top{\tilde{\mathtt{t}}}$, where $A$ is a matrix of the right singular vectors of a normalized Genotype matrix, that correspond to the top $10$ singular values~\cite{price2006principal}.
The outcome model is selected from logistic Lasso linear models with various regularization strengths, via cross validation within the training data ($60\%$ of the dataset).
We defer details about the experimental setup to~\cref{appsec:exps}.
We then use this outcome model in \gls{ours} to compute causal effects on the whole filtered dataset.
The effects are computed one \gls{snp} at a time.
First, for each person ${\tilde{\mathtt{t}}}$, create ${\tilde{\mathtt{t}}}_i^1, {\tilde{\mathtt{t}}}_i^0$ which correspond to the $i$th \gls{snp} set to $1$ and $0$ respectively, with all other \glspl{snp} same as ${\tilde{\mathtt{t}}}$.
Randomly sample a $h(\mathtt{t}^*_2)$ from the marginal $p(h(\mb{t}))$ and, using the outcome model $P_\theta$, compute $\phi({\tilde{\mathtt{t}}}, i) = \log \nicefrac{P_\theta(y = 1\mid \mathtt{t}^\prime({\tilde{\mathtt{t}}}_i^1, h(\mathtt{t}^*_2)))}{P_\theta(y = 1\mid \mathtt{t}^\prime({\tilde{\mathtt{t}}}_i^0, h(\mathtt{t}^*_2)))}$.
The average effect of \gls{snp} $i$ is obtained by averaging across all persons: $\sum_{{\tilde{\mathtt{t}}}}\nicefrac{\phi({\tilde{\mathtt{t}}},i)}{N}$.
Any \gls{snp} that beats a specified threshold of effect is deemed relevant to Celiac disease by \gls{ours}.
We use a $60-40\%$ train-test split, and outcome model selection is done via cross-validation within the training set.
We did $5$-fold cross-validation using just the training set.
We use Scikit-learn~\cite{pedregosa2011scikit} to fit the outcome models and for cross-validation.
\paragraph{Results}
The best outcome model was a Lasso model, trained with regularization constant $10$.
We select relevant \glspl{snp} by thresholding estimated effects at a magnitude $>0.1$.
From $1050$ \glspl{snp} ($1000$ not reported before) \gls{ours} returned $31$ \glspl{snp}, out of which $13$ were previously reported as being associated with Celiac disease~\citep{dubois2010multiple,sollid2002coeliac,hunt2008novel,adamovic2008association}.
In~\cref{appsec:real-exp} we plot the true positive and false negative rates of identifying previously reported~\glspl{snp}, as a function of the effect threshold.
\begin{wraptable}[14]{r}{0.43\textwidth}
\centering
\begin{tabular}{ c c c }
\toprule
\gls{snp} & \textsc{Effect.} & \textsc{Coef.} \\
\midrule
rs13151961 & $0.17$ & $0.32$ \\
rs2237236 & $0.17$ & $0.00$ \\ \midrule
rs1738074 & $-0.16$ & $-0.23$ \\
rs11221332 & $-0.15$ & $-0.24$ \\
\bottomrule
\end{tabular}
\caption{A few \glspl{snp} previously reported as relevant and recovered by \gls{ours}, with
estimated effects and Lasso coefficients.
\gls{ours} produces effect estimates that do not rely purely on the coefficients.
}
\label{tab:snps-est}
\end{wraptable}
In~\cref{tab:snps-est}, we list a few \glspl{snp} that were both deemed relevant by \gls{ours} and were reported in existing literature~\citep{dubois2010multiple,sollid2002coeliac,hunt2008novel,adamovic2008association}, their effects, and their Lasso coefficients.
The full list is in~\cref{apptab:snps-est} in~\cref{appsec:exps}.
If \gls{ours} cannot adjust for confounding, the Lasso coefficients would dictate the effects; $0$ coefficient means $0$ effect.
However, the two pairs of \glspl{snp} in \cref{tab:snps-est} show that the effects estimated by \gls{ours} do not rely solely on the Lasso coefficients.
For the first pair (rs13151961, rs2237236), the effect is the same but the coefficient of one is $0$, while the other is positive.
We note that rs2237236 was found to be associated with ulcerative colitis~\citep{hindorff2009potential,anderson2011meta}, which
is an inflammatory bowel disease that has been reported to share some common genetic basis with celiac disease~\citep{pascual2014inflammatory}.
For the second pair, (rs1738074, rs11221332), the magnitude of the effect is smaller for the former, but the coefficient is larger.
Thus, \gls{ours} adjusts for confounding factors that the outcome model ignored. |
Subsets and Splits